Metal-Oxo and Metal-Peroxo Species in Catalytic Oxidations [1 ed.] 3540669434, 9783540669432

The present volume is a description of the current knowledge on the different metal-oxo and metal-peroxo species which a

453 45 5MB

English Pages 323 [327] Year 2000

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Metal-Oxo and Metal-Peroxo Species in Catalytic Oxidations  [1 ed.]
 3540669434, 9783540669432

Table of contents :
Cover......Page 1
Structure and Bonding 97......Page 2
Metal-Oxo and Metal-Peroxo Species in Catalytic Oxidations......Page 4
ISBN 3-540-66943-4......Page 5
Editors......Page 6
Preface......Page 7
Contents......Page 9
001-035......Page 11
037-060......Page 46
061-089......Page 70
091-123......Page 99
125-144......Page 132
145-177......Page 152
179-211......Page 185
213-236......Page 218
237-285......Page 242
287-303......Page 291
Author Index Volumes 1-97......Page 308

Citation preview

Structure and Bonding

Editorial Board: A.J. Bard. I.G. Dance P. Day J.A. Ibers T. Kunitake T.J. Meyer D.M.P. Mingos H.W. Roesky J.-P. Sauvage A. Simon. F. Wudl

.

Springer Berlin Heidelberg New York Barcelona Hong Kong London Milan Paris Singapore Tokyo

Metal-0x0 and Metal-Peroxo Species in Catalytic Oxidations Volume Editor: B. Meunier

With contributions by W. Adam, F. Banse, J. Bernadou, A.G. Blackman, R.M. Burger, R.H. Crabtree, H. Fujii, J.-J. Girerd, W.A. Herrmann, F.E. Kiihn, B. Meunier, C.M. Mitchell, C.R. Saha-Moller, D. Schroder, H. Schwarz, S. Shaik, P.E.M. Siegbahn, A.J. Simaan, W.B. Tolman, J.S. Valentine, Y. Watanabe, 0.Weichhold, D.L. Wertz

Springer

The series Structure and Bonding publishes critical reviews on topics of research concerned with chemical structure and bonding. The scope of the series spans the entire Periodic Table. It focuses attention on new and develovinp. areas of modern structural and theoretical chemistry such as nanostructures, molecular electronics, designed molecular solids, surfaces, metal clusters and supramolecular structures. ~ h ~ &and i l spectroscopic techniques used to determine, examine and model structures fall within the puniew of Structure and Bonding to the extent that the focus is on the scientific results obtained and not on specialist information concerning the techniques themselves. Issues associated with the development of bonding models and generalizations that illuminate the reactivity pathways and rates of chemical processes are also relevant.

.-

As a rule, contributions are specially commissioned. The editors and publishers will, however, always be pleased to receive suggestions and supplementary information. Papers are accepted for Structure and Bonding in English. In references Structure and Bonding is abbreviated Struct.Bond. and is cited as a journal. Springer WWW home page: http:llwww.springer.de

ISSN 0081-5993 ISBN 3-540-66943-4 Springer-Verlag Berlin Heidelberg New York CIP Data applied for This work is subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965, in its current version, and permission for use must always be obtained from Springer-Verlag. Violations are liable for prosecution under the German Copyright Law. Springer-Verlag is a company in the BertelsmannSpringer publishing group 0 Springer-Verlag Berlin Heidelberg 2000 Printed in Germany The use of registered names, trademarks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. Typesetting: Scientific Publishing Senices (P) Ltd, Madras Production editor: Christiane Messerschmidt, Rheinau Cover: Medio V. Leins, Berlin Printed on acid-free paper SPIN: 10702329 0213020 - 5 4 3 2 1 0

Volume Editor Professor Dr. Bernard Meunier Laboratoire d e Chimie d e Coordination d u CNRS 205 route d e Narbonne 31077 Toulouse Cedex 4, France E-mail: [email protected]

Editorial Board Prof. Allen J. Bard Department of Chemistry and Biochemistry University of Texas 24th Street and Speedway Austin, Texas 78712, USA E-maiE [email protected]

Prof. Ian G. Dance Department of Inorganic and Nuclear Chemistry School of Chemistry University of New South Wales Sydney, NSW 2052, Australia E-mail: [email protected]

Prof. Peter Day, FRS Director and Fullerian Professor of Chemistry The Royal Institution of Great Britain 21 Albemarle Street London WIX 4BS, UK E-mail: [email protected]

Prof. James A. Ibers Department of Chemistry North Western University 2145 Sheridan Road Evanston, Illinois 60208-3113, USA E-mail: [email protected]

Prof. Toyohi Kunitake Faculty of Engineering: Department of Organic Synthesis Kyushu University Hakozaki 6-10-1, Higashi-ku Fukuoka 812, lapan E-mail: [email protected]

Prof. Thomas J. Meyer

Prof. D. Michael P. Mingos Principal St. Edmund Hall Oxford OX1 4AR, UK E-mail: [email protected] Prof. Jean-Pierre Sauvage Facult; de Chimie Laboratories de Chimie Organo-Minerale Universitk Louis Pasteur 4, rue Blaise Pascal 67070 Strasbourg Cedex, France E-mail: [email protected] Prof. Fred Wudl Department of Chemistry University of California LosAngeles, CA 90024-1569, USA E-mail: [email protected]

Department of Chemistry University of North Carolina at Chapel Hill Venable and Kenan Laboratory CB 3290 Chapel Hill, North Carolina 27599-3290, USA E-mail: [email protected] Prof. Herbert W. Roesky Institut fiir Anorganische Chemie der Universitat Gottingen Tammannstrage 4 D-37077 Gottingen, Germany E-mail: [email protected] Prof. Arndt Simon Max-Planck-Institut fiir Festkijrperforschung Heisenbergstrage 1 70569 Stuttgart, Germany E-mail: [email protected]

Preface

Free radical oxidations of hydrocarbons in the liquid phase are still the only large-scale method for obtaining most of the oxygen-containing chemical raw materials. These autooxidation reactions involve the reaction of the triplet state of dioxygen with a carbon-centered radical as elementary step. For the last twenty years, many research groups have worked in the ®eld of transition-metal catalyzed oxidations having in mind two main challenges: (i) the understanding of the mechanism of oxidation reactions catalyzed by mono- and di-oxygenases at the molecular level (how to hydroxylate an aliphatic CAH bond at room temperature with retention of con®guration?) and (ii) the use of the knowledge accumulated on oxygenases with biomimetic models to create new metal-catalyzed oxidation methods as selective as possible, without free radical intermediates. Probably more than 2000 articles have been published on metal-catalyzed oxidations at ambient temperature between 1980 and 1999. The present volume is not an attempt to produce exhaustive reviews on all these catalytic oxidations performed in mild conditions (the so-called biomimetic or bioinspired oxidations), but is a selection of different contributions to describe the current knowledge on the different metal-oxo and metal-peroxo species which are involved in catalytic oxidations. These highly reactivity metal-oxygen species are the key intermediates in heme- or non-heme oxygenases, as well as in metal-catalyzed oxidations. The accumulated knowledge on these reactive entities will facilitate a better understanding of the different intermediates involved in the catalytic cycles of enzymes like cytochromes P450, non-heme oxygenases, copper oxygenases, in the oxidation of water by the manganese-center of Phyotosystem II, in the chemistry of dioxiranes and also in DNA cleavage performed by activated bleomycin. The description of these metal-oxo and metal-peroxo species is based on physicochemical data, reactivity results and also on theoretical calculations. The titles of the ten different contributions are as follows: 1. Active iron-oxo and iron-peroxo species in cytochrome P-450 and peroxidases; oxo-hydroxo tautomerism with water-soluble metalloporphyrins (Bernard Meunier and Jean Bernadou), 2. Nucleophilicity of iron-peroxo porphyrin complexes. (Joan S. Valentine and Diana Wertz), 3. Characterization of high-valent oxo-metalloporphyrins. (Yoshihito Watanabe and Hiroshi Fujii), 4. Characterization, orbital description, and reactivity patterns of transition-metal oxo species in the gas phase (Detfel SchroÈder, Helmut

VIII

Preface

Schwarz and Sason Shaik), 5. Theoritical aspects on high-valent metal-oxo or -peroxo species in methane monooxygenases and the manganese-center in Photosystem II. (P.E.M. Siegbahn and R.C. Crabtree), 6. Characterization and properties of non-heme iron-peroxo complexes (Jean-Jacques Girerd, FreÂdeÂric Banse and Ariane J. Simaan), 7. Copper-dioxygen and copper-oxo species relevant to copper oxygenases and oxidases (William B. Tolman and Allan Blackman), 8. Rhenium-oxo and rhenium-peroxo complexes in catalytic oxidations (Fritz E. KuÈhn and Wolfgang A. Herrmann), 9. Structure, reactivity and selectivity of metal-peroxo complexes versus dioxiranes (Waldemar Adam, Catherine M. Mitchell, Chantu R. Saha-Moeller and Olivier Weichold) and 10. Nature of Activated Bleomycin (Richard M. Burger). We hope that these different contributions on metal-oxo and metal-peroxo species will help the readers in their efforts to better understand the mechanistic aspects of metal-catalyzed oxidations. This volume illustrates new directions for chemists engaged in the discovery of new ef®cient and selective oxidation catalysts, molecular biologists interested by enzyme-catalyzed oxidations and pharmacologists working on the development of potential antitumoral or antiviral drugs able to degrade the nuclei acids of tumor cells or viruses. As Editor of this volume, I am deeply grateful to the authors of the different contributions. They all agreed with enthusiasm to participate in the preparation of the present Volume despite the pressure which is on the shoulders of talented researchers. Toulouse, March 2000

Bernard Meunier

Contents

Active Iron-0x0 and Iron-Peroxo Species in Cytochromes P450 and Peroxidases; 0x0-Hydroxo Tautomerism with Water-Soluble Metalloporphyrins B. Meunier, J. Bernadou. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nucleophilicity of Iron-Peroxo Porphyrin Complexes D.L. Wertz, J.S. Valentine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Characterization of High-Valent 0x0-Metalloporphyrins Y. Watanabe, H. Fujii.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Characterization, Orbital Description, and Reactivity Patterns of Transition-Metal 0x0 Species in the Gas Phase D. Schroder, H. Schwarz, S. Shaik. . . . . . . . . . . . . . . . . . . . . . Quantum Chemical Studies on Metal-0x0 Species Related to the Mechanisms of Methane Monooxygenase and Photosynthetic Oxygen Evolution P.E.M. Siegbahn, R.H. Crabtree. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Characterization and Properties of Non-Heme Iron Peroxo Complexes 1.-J. Girerd, F. Banse, A.J. Simaan . . . . . . . . . . . . . . . . . . . . . . . . . . . . Copper-Dioxygen and Copper-0x0 Species Relevant to Copper Oxygenases and Oxidases A.G. Blackman, W.B. Tolman . . . . . . . . . . . . . . . . . . . . . . Rhenium-0x0 and Rhenium-Peroxo Complexes in Catalytic Oxidations F.E. Kiihn, W.A. Herrmann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes W. Adam, C.M. Mitchell, C.R. Saha-Moller, 0.Weichhold . . . . . . . . . . Nature of Activated Bleomycin R.M.Burger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Author Index Volumes 1-97

...............................

305

Contents of Volume 93 Bonding and Charge Distribution in Polyoxometalates: A Bond Valence Approach Volume Editor: D.M.P. Mingos Bond Length-Bond Valence Relationships with Particular Reference to Polyoxometalate Chemistry K.H. Tytko, J. Mehmke, D. Kurad A Bond Model for Polyoxometalate Ions Composed of MO, Octahedra (MO, Polyhedra with k > 4) K.H. Tyko

Boading and Charge Distribution in Isopolyoxometalate Ions and Relevant Oxides - A Bond Valence Approach K.H. Tytko, J. Mehmke, S. Fischer

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases; Oxo-Hydroxo Tautomerism with Water-Soluble Metalloporphyrins Bernard Meunier1, Jean Bernadou2 Laboratoire de Chimie de Coordination du CNRS, 205 route de Narbonne, 31077 Toulouse cedex 4, France 1 E-mail: [email protected] 2 E-mail: [email protected] Heme-containing monooxygenases are able to catalyze two different classes of oxidation reactions. The ®rst class includes oxygenation reactions (hydroxylation, epoxidation, N- or S-oxide formation, etc.) which are mediated by an electrophilic oxidative species. The second class is represented by the oxidative deformylation of aldhehydes and involves a nucleophilic oxidant as active intermediate. The reductive activation of molecular oxygen by cytochromes P450 generates a nucleophilic iron(III)-peroxo species which produces by protonation an electrophilic high-valent iron-oxo [formally an iron(V)oxo] responsible for electrophilic oxygen atom transfers. The nucleophilic properties of the iron(III)-peroxo intermediate in cytochrome P450 are due to the porphyrin ring acting as electron reservoir and also to the negative charge accumulated on the proximal cysteine during the initial reduction step of the catalytic cycle. The nature of the high-valent iron-oxo species generated in the catalytic cycle of heme-peroxidases will be also discussed. Among the different methods for studying the oxygenation reactions mediated by high-valent metal-oxo porphyrin complexes, the recent discovery of the ``oxo-hydroxo tautomerism'' provides a useful tool to investigate the mechanism of O-atom transfer reactions in aqueous media. Keywords: Metal-oxo, Metal-peroxo, Electrophilic, Nucleophilic, Cytochrome, Peroxidase,

Oxo-hydroxo tautomerism, Metalloporphyrin

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3

2

Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4

2.1

Structural Aspects of Cytochromes P450 and Substrate Binding . . . . . . . . . . . . . . . . . . . . . . . . . . . Catalytic Cycle of Cytochromes P450 . . . . . . . . . . . . . . . . First Reduction Step . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxygen Binding Step . . . . . . . . . . . . . . . . . . . . . . . . . . . . Transfer of the Second Electron (Formation of a Nucleophilic Iron-Peroxo Species) . . . . . . First Protonation Step (Formation of a Nucleophilic Iron-Hydroperoxo Species) . . . . . . . . . . . . . . . . . . . . . . . Second Protonation Step (Generation of an Electrophilic Iron-Oxo Species) . . . . . . . Reactivity of Iron(III)-Peroxo Porphyrin Complexes . . . . . Characterization and Reactivity of High-Valent Metal-Oxo Porphyrin Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2 2.2.1 2.2.2 2.2.3 2.2.4 2.2.5 2.3 2.4

. . . .

. . . .

. . . .

. . . .

4 7 7 8

....

8

....

10

.... ....

12 16

....

17

Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

2

B. Meunier á J. Bernadou

2.5 2.5.1 2.5.2

Active Species in Heme-Peroxidase Catalytic Cycles . . . . . . . . . Horseradish Peroxidase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chloroperoxidase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

18 19 20

3

Oxo-Hydroxo Tautomerism with Water-Soluble Metalloporphyrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

22

3.1

3.4.5 3.4.6

A New Type of Tautomerism: The Oxo-Hydroxo Tautomerism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Some Previous Data on O-Exchange of Metal-Oxo Species with Bulk Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxo-Hydroxo Tautomerism Observed in Different Metalloporphyrin-Catalyzed Oxygenation Reactions . . . . . . . Epoxidation Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hydroxylation Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . Quinone Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Required Conditions to Observe an Oxo-Hydroxo Tautomerism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . An Axial Ligand Competitive to the Hydroxo Ligand Inhibits Oxo-Hydroxo Tautomerism . . . . . . . . . . . . . . . . . . . . . . . . . A Competitive Autoxidation Route Lowers Incorporation of Oxygen from Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . Temperature Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Differences in Exchange Kinetics for Metal(IV)-Oxo and Metal(V)-Oxo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Role of the Ratio Water/Substrate Concentrations . . . . . . . . . Other Parameters (pH, etc.) . . . . . . . . . . . . . . . . . . . . . . . . .

..

23

..

23

. . . .

. . . .

24 24 26 27

..

28

..

28

.. ..

29 29

.. .. ..

30 31 31

4

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

32

5

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

32

3.2 3.3 3.3.1 3.3.2 3.3.3 3.4 3.4.1 3.4.2 3.4.3 3.4.4

List of Abbreviations CBZ Cpd I or II CPO DFT ee F20TPP KIE HRP L m-CPBA

carbamazepine Compound I or II of a heme-peroxidase, an iron(IV)-oxo radical-cation species, or an iron(IV)-oxo species, respectively chloroperoxidase density functional theory enantiomeric excess dianion of the meso-tetrakis(penta¯uorophenyl)porphyrin ligand kinetic isotope effect horseradish peroxidase a neutral ligand (e.g., pyridine) providing two electrons to a metal center via a donor-acceptor bond meta-chloroperobenzoic acid

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

5-MF NADH PPIX TDCPPS TF4TMAP TMP TMPyP X

3

5-methylene-2-furanone (NADPH) reduced form of nicotinamide adenine dinucleotide (phosphate) protoporphyrin-IX dianion of the meso-tetrakis(2,6-dichloro-3-sulfonatophenyl)porphyrin ligand dianion of meso-tetrakis(2,3,5,6-tetra¯uoro-4-N,N,N-trimethylaniliniumyl)porphyrin ligand the dianion of meso-tetramesitylporphyrin ligand the dianion of meso-tetrakis(4-methylpyridiniumyl)porphyrin ligand an anionic ligand (halide, deprotonated cysteine RS), superoxide anion O2 , monoanion of hydrogen peroxide HOO), . . . etc.)

1 Introduction The cytochromes P450 constitute a large family of cysteinato-heme enzymes (over 500 members) present in all forms of life (plants, bacteria, mammals) and they play a key role in the oxidative transformation of endogeneous and exogenous molecules [1±3]. These enzymes are monooxygenases able to catalyze the insertion of one oxygen atom of dioxygen within many different substrates, the second oxygen atom of O2 being reduced to a water molecule. The two electrons of this reductive activation of molecular oxygen are provided by NAD(P)H via a reductase. These enzymes are located in the membrane of the endoplasmic reticulum and are able to perform various dif®cult oxygenation reactions. Cytochromes P450 catalyze the hydroxylation of saturated carbon-hydrogen bonds (Eq. 1), the epoxidation of double bonds, the oxidation of heteroatoms, dealkylation reactions, oxidations of aromatics, ... etc.: Cytochrome P450

RH ‡ O2 ‡ 2e ‡ 2H‡ ƒƒƒƒƒƒƒƒƒƒƒƒƒƒ! ROH ‡ H2 O

…1†

Being a triplet (two unpaired electrons in the ground state), molecular oxygen is unreactive towards organic molecules at low temperatures. The reaction of dioxygen with the single state of organic substrates is spinforbidden [4]. Consequently, the oxygenation of organic molecules at physiological temperatures must involve the modi®cation of the electronic structure of one of the partners. Living systems mainly use enzymes to modify the electronic structure of dioxygen to a form which is adapted for the desired oxidation reaction. This modi®cation can be performed by metal-dependent oxygenases, like cytochromes P450 or non-heme metalloenzymes (e.g., methane monooxygenase), or by ¯avin-containing enzymes which do not possess metal-based prosthetic groups. Forty years after the isolation and the characterization of cytochromes P450 [5], the exact nature of the active species responsible for the oxygen insertion step is still a matter of debate. After an

4

B. Meunier á J. Bernadou

``iron-oxenoid'' period during the 1970s [5] came the ``high-valent iron-oxo'' period, mainly based on elegant works using single oxygen atom donors with cytochrome P450 itself and with chemical models employing synthetic metalloporphyrins [3, 6±10]. More recently, the hypothesis of an electrophilic ``iron(III)-hydroperoxo'' appeared in the literature [11], partly inspired by oxidations catalyzed by non-heme iron complexes and DNA sugar oxidation performed by activated iron-bleomycin within tumor cells. In Sect. 2 we will discuss the nature of the different species generated during the catalytic cycle of cytochromes P450 and heme-peroxidases taking into consideration the knowledge which has been accumulated on these enzymes themselves and from biomimetic oxidations over the past two decades [12]. In Sect. 3 we will then present a survey of recent observations of an ``oxohydroxo tautomerism'' that has been collected on oxygenation reactions catalyzed by water-soluble metalloporphyrin complexes [13]. This oxohydroxo tautomerism is a useful tool which can be employed in mechanistic studies of oxygen atom transfer performed by high-valent metal-oxo species in aqueous solutions.

2 Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases Before entering into the details of the catalytic cycle of cytochromes P450, we will ®rst report the essential structural features of this fascinating class of monooxygenases. Subsequently, the main steps of the catalytic cycle of hemeperoxidases will be presented. 2.1 Structural Aspects of Cytochromes P450 and Substrate Binding

The name cytochrome P450 arises from the fact that the reduced protein ef®ciently binds carbon monoxide with a strong absorption band (the Soret p±p* band) of the PPIX-Fe(II)-CO chromophore at 450 nm. This spectroscopic property has been very useful since the early studies on cytochromes P450 to monitor the presence of this enzyme in the microsomal fraction of liver tissues [3]. Cytochrome P450 is no longer a black box. In 1986, Poulos et al. provided the ®rst tri-dimensional structure of the cytochrome P450cam of Pseudomonas putida (Scheme 1) which catalyzes the stereospeci®c hydroxylation of the exo C5-H bond of camphor (Scheme 2) [14, 15]. This microorganism has the ability to use this terpene as its only source of carbon and energy, camphor being hydroxylated and then metabolized to isobutyrate and acetate. The electrons which are necessary at different steps of the catalytic cycle are provided by NADH-coupled ¯avoprotein (putidaredoxin reductase) via an iron-sulfur protein (putidaredoxin). P450cam is a 45,000 Dalton polypeptide chain containing a single ferric protoporphyrin-IX and a cysteine, Cys-357, as axial ligand (a proximal cysteinato ligand is also present in chloroperoxidase).

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

5

Scheme 1. Structure of the cytochrome P450cam of Pseudomonas putida. Reproduced with

permission from [14]

Scheme 2. Stereospeci®c hydroxylation of the exo CAH bond at position 5 of camphor by

cytochrome P450cam

As in many other heme proteins, the FeIII state of the resting enzyme equilibrates between the low spin (S = 1/2) and the high spin state (S = 5/2), but the low spin state is favored in the absence of substrate and the sixth position of the octahedron is occupied by a water molecule (an aqua ligand). In the presence of camphor, the spin equilibrium of the iron(III) center is shifted toward the high spin form and the axial water molecule is removed, as well as the other water molecules located within the rather hydrophobic pocket of the active site of this monooxygenase. This modi®cation of the coordination sphere of the metal center induces a change in the redox potential of the iron center, shifting from )300 to )170 mV after substrate binding, which facilitates the reduction of the pentacoordinated ferric center by the reductase in the next step of the catalytic cycle.

6

B. Meunier á J. Bernadou

The kinetic and thermodynamic parameters of the reversible binding of camphor in cytochrome P450cam have been well studied by Grif®n and Peterson [16]. The association of camphor is a second order reaction with a rate constant of 4.1 ´ 106 M)1 s)1, while the dissociation is a slow ®rst order reaction with a rate constant of 6.0 s)1. The camphor binding corresponds to an entropy change (DS) of 26 cal mol)1 K)1 and a free-energy change (DG) of )7.7 kcal mol)1 at 21 °C. The substrate binding is an entropy-driven process, as the favorable entropy change associated with the removal of water molecules from the hydrophobic active site by camphor binding represents the major component of the overall binding energy [16]. The catalytic cycle of cytochrome P450 is triggered by the entry of the substrate into the active site displacing the axial water molecule (Scheme 3). Consequently the iron is more displaced from the plane of the porphyrin ring: Ê compared to 0.30 A Ê in the resting state of the enzyme. In addition, Tyr0.44 A 96 which binds with the keto group of camphor acts as a probe for the polarity changes within the active site of the enzyme when camphor binds. The sensitivity of Tyr-96 to environmental polarity favors the access of bulk water molecules in the product-enzyme complex formed with 5-exo-hydroxycamphor, thus facilitating the product release from the enzyme [17]. X-ray diffraction studies on P450cam crystallized with and without the camphor substrate show the existence of a hydrogen bond interaction between the 2-keto group of camphor and Tyr-96 in addition to weak enzyme hydrophobic interactions with Phe-87, Leu-244, Val-247, Thr-252, and Val295. All these different low-energy interactions are suf®cient to dictate the exact ®tting of the substrate inside the active site, namely putting the 5-exo

Scheme 3. Schematic representation of the different possible intermediates in the catalytic

cycle of cytochrome P450

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

7

hydrogen atom close to the putative active iron(V)-oxo species, then making possible the speci®c hydroxylation at the 5-exo position. This speci®city is modi®ed when using analogs of camphor [18]. The crystal structures of P450terp (an a-terpineol monooxygenase), P450eryF (a monooxygenase involved in the biosynthesis of erythromycin), and P450BM3 (a fatty acid hydroxylase) have been determined [18, 19]. In all these bacterial enzymes, a threonine residue is present in the active site to deliver protons to the iron-dioxygen intermediate during catalysis. Recently, a structure of P450cam modi®ed with two electroactive ferrocenyl groups has been solved [20]. 2.2 Catalytic Cycle of Cytochromes P450

2.2.1 First Reduction Step After the substrate binding step discussed above (A ® B in Scheme 3), the next step is the reduction of the iron(III) center to a ferrous state. A large positively charged area on the outside of P450s is involved in the interaction of the oxygenase with the reductase, suggesting that the contact between both partners is primarily electrostatic [19]. The rate of the ®rst reduction step is relatively slow (k = 35 s)1). The reduced form of cytochrome P450 (intermediate C in Scheme 3) is an extremely ef®cient reducing agent. For example, polyhalogenated hydrocarbons (halothane, carbon tetrachloride, etc.) produce a stable iron(II)-carbene species via a reductive dehalogenation [3]. The nature of the sulfur-iron bond is probably modi®ed in the ferrous intermediate C compared to the starting ferric derivative A. As a matter of fact, Ê in the ferric P450cam (A, the sulfur-iron bond distance changes from 2.20 A Ê in the CO-adduct of intermediate C with the ferrous Scheme 3) to 2.35±2.41 A Ê from the porphyrin plane [21]. This important ion only displaced by 0.02 A structural change has been discussed in terms of coordination chemistry and the possible modi®cation of the nature of the sulfur-iron bond. Despite the high resolution of the X-ray structure, the authors preferred to attribute this distance modi®cation to experimental errors rather than to a possible modi®cation of the nature of the iron-sulfur bond. The reduction of B, a L2X3 ferric complex, would formally lead to a negatively charged L2X3 ferrous complex or to a L3X2 ferrous complex with a neutral cysteine ligand (RSH, i.e., an L ligand) instead of a cysteinato ligand (RS-, an X ligand). The protonation of the proximal cysteine would give rise to a cytochrome having a Soret band at 420 nm, excluding the hypothesis of a full protonation of the sulfur atom, but theoretical calculations [22] and data on models systems [23] demonstrate that a negative charge exists on intermediate C (this charge is delocalized on the whole prosthetic group, including the sulfur atom, and is certainly playing a key role in the heterolytic cleavage of the OAO bond of the ferrichydroperoxo complex leading to the high-valent iron-oxo species. The cysteinyl sulfur atom is hydrogen bonded to the amide proton of three amino

8

B. Meunier á J. Bernadou

acids in P450cam ± see [18, 24]). This charge is not explicitly indicated on intermediates C to E in Scheme 2, like in any other text-book on cytochrome P450, for simpli®cation, but plays a key role in the electron density of the irondioxygen intermediate. The representation of this negative charge will be also discussed later (see Scheme 11). The reduced ferrous cytochrome P450cam has a d6 high spin state. The next step is the binding of dioxygen to the ferrous state of this metalloenzyme. 2.2.2 Oxygen Binding Step Triplet dioxygen reacts with ferrous cytochrome P450cam with a second order rate constant of 1.7 ´ 106 M)1 s)1 to produce a stable dioxygen adduct (Kaff = 106 M)1) [2]. One electron from the iron center and one from triplet oxygen are pairing to create an iron(III)-oxygen bond. This oxygen-iron complex (compound D in Scheme 3) is relatively stable, but can dissociate to an iron(III) and superoxide anion with a rate constant of 0.01 s)1 at room temperature. Within the enzyme, the release of superoxide is followed by its disproportionation and generates hydrogen peroxide, a source of harmful hydroxyl radicals (such a step is called a ``decoupling reaction'' in the vocabulary of P450 chemistry). So, the intermediate D can be regarded as an g1-superoxide ion coordinated to an iron(III) center with an unpaired electron on the terminal oxygen atom. An OAO stretching vibration has been observed at 1141 cm)1 in the resonance Raman spectrum of P450cam under catalytic conditions [25]. The Fe-OAO bending mode has been observed at 401 cm)1 by resonance Raman spectroscopy with an angle of approximately 125±130° [26]. The electronic structure and the chemical properties of this ferrousdioxygen intermediate of P450 enzymes are different from the analogue stable ferrous-dioxygen states of molecular oxygen carriers such as hemoglobin and myoglobin. In these heme-dioxygen carriers, the reduced ferrous state is neutral with the proximal position being occupied by a histidine nitrogen atom. The basicity of the proximal ligand increases the stability of the dioxygen-iron adduct [27]. In this respect, a proximal cysteine is not as good as a histidine ligand to stabilize a FeII-heme-dioxygen adduct. 2.2.3 Transfer of the Second Electron (Formation of a Nucleophilic Iron-Peroxo Species) The second reduction step is the rate determining step in all different cytochrome P450s. This relatively slow step (k = 17 s)1 in cytochrome P450cam) generates a negatively charged iron(III)-peroxo complex (intermediate E in Scheme 3) which is probably quickly protonated at this stage. This intermediate with a red-shifted band at 350±450 nm with a split Soret band has been observed by UV-visible spectroscopy in a D251N P450cam mutant (the aspartic residue at 251 being replaced by an asparagine) [28, 29]. The aspartic residue plays a key role in the kinetics of the proton transfer to this

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

9

intermediate E. In the mutant Asp251Asn, the rate-determining step is not the reduction step, but the proton transfer to generate an FeAOOH entity (this protonated form of complex E is not depicted in Scheme 3). The turnover rate for camphor hydroxylation of this mutant is 1±2 orders of magnitude slower than native P450cam [29]. This non-protonated iron(III)peroxo complex E is an obvious candidate to explain some particular reactions catalyzed by cytochrome P450s which obviously involve a nucleophilic oxidant intermediate. This is the case with both the oxidative decarbonylation of aldehydes and the ®nal step of the aromatization of the cycle A of androstenedione in the biosynthesis of estrone. The intermediate E is certainly a stronger nucleophile than the corresponding protonated form FeIIIAOOH. However, both entities can be involved in these nucleophilic oxidations depending on the relative rates of the nucleophilic addition of the non-protonated iron-peroxo and of its protonation. If the proton transfer is faster than the nucleophilic attack, then only FeIIIAOOH is responsible for the nucleophilic oxidations observed with some substrates. Aldehydes are usually oxidized to carboxylic acids by cytochrome P450s, but a second reaction pathway has been identi®ed with some particular substrates. Cyclohexene is formed during the oxidation of cyclohexanecarboxaldehyde catalyzed by P4502B4, a liver P450 induced by phenobarbital (Scheme 4) [30]. This reaction was not observed when iodosylbenzene was used as single oxygen atom source, indicating clearly that the high-valent ironoxo intermediate was not involved in this reaction. The same oxidative deformylation has been observed with other aldehydes including citronellal [31]. The carbon atom of the aldehyde function is eliminated as formic acid. The initial nucleophilic attack of the iron-peroxo entity on the carbonyl function is probably followed by a concerted reaction involving 6-electron like in a Grob fragmentation (see Scheme 4 for a description of the reaction pathways). A second classical example of the role of a nucleophilic iron-peroxo species in a P450 cycle is the third step of the aromatization reaction of the A-ring of androst-4-ene-3,17-dione to estrone catalyzed by human placental aromatase [32, 33]. This reaction involves three consecutive oxidative steps: two hydroxylations at the 19-methyl group and a ®nal oxidative decarbonylation of the intermediate 19-aldehyde. All three reactions are stereospeci®c: the ®rst

Scheme 4. Oxidative decarbonylation of cyclohexanecarboxaldehyde to cyclohexene and

formic acid catalyzed by cytochrome P450

10

B. Meunier á J. Bernadou

hydroxylation at the 19-methyl occurs with retention of con®guration, the second removes the 19-pro-R hydrogen yielding a gem-diol which produces the 19-aldehyde derivative (see Scheme 5). The ®nal step is the nucleophilic attack of the iron-peroxo intermediate generated during the third catalytic cycle (the substrate remains effectively within the active site for three cycles). This decarbonylation reaction also involves the stereospeci®c removal of the 1b and the 2b hydrogens to produce the phenolic A-ring of estrone while the 19-methyl group is eliminated as formic acid. This type of nucleophilic oxidation of an iron-peroxo onto a carbonyl group has also been evidenced in the ®nal step of the formation of NO during the oxidative degradation of arginine catalyzed by NO synthases, a family of hemeenzymes able to produce nitric oxide in vivo [34]. However, it should be mentioned that the oxidation of heme to biliverdin by heme oxygenase might be due to an electrophilic iron(III)-peroxo species [35]. In general, the main reactions catalyzed by P450 enzymes are the incorporation of an oxygen atom into a substrate: hydroxylation of an aliphatic CAH bond, epoxidation of an ole®n, oxidation of an heteroatom, etc. All these oxidations involve an electrophilic metal-oxo species as we will see below. 2.2.4 First Protonation Step (Formation of a Nucleophilic Iron-Hydroperoxo Species) Threonine-252 in P450cam plays a key role in the protonation of the ironperoxo intermediate (see Scheme 6). This threonine residue is highly conserved among cytochrome P450s.

Scheme 5. Mechanism of the aromatization of the A-ring of androst-4-ene-3,17-dione to

estrone catalyzed by human placental aromatase

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

11

Scheme 6. Essential role of Thr-252 and Asp-251 residues in proton delivery to the iron-

dioxygen species in cytochrome P450s

The ®rst protonation produces an FeIIIAOOH intermediate which can behave as a nucleophile (see above for the discussion on the reduced nucleophilicity of this protonated form compared to the corresponding non-protonated species, intermediate E in Scheme 3). The close interaction of Thr-252 with Asp-251 allows fast proton transfers from protonated forms of Lys-178 and Arg-186. When Thr-252 is replaced in P450cam by an alanine, the mutant enzyme is producing mainly hydrogen peroxide and water molecules in large excess compared to the expected camphor hydroxylation (the socalled uncoupled reactions) [36, 37]. The essential role of the Asp-251 in the proton transfer to the FeIIIAOAO) intermediate (E in Scheme 3) has been studied in cytochrome P450cam using site-directed mutagenesis, crystallography, and kinetic solvent isotope effects [29]. The turnover rates of the Asp251Asn mutant in various proton-deuterium mixtures have been determined. The kinetic solvent isotope effect is larger, with a value of 10 compared to 1.8 for the wild-type enzyme. The hydrogen bond network created with the aspartic residue is broken in the mutant and the dioxygen adduct is more open to bulk water molecules. (Side-commentary: it should be noted that the oxidation states of the iron centers of intermediates 4a and 4b in Fig. 1 of [29] are not correct: the protonation of the terminal oxygen atom is not going to change the oxidation state of the iron center from II to III. In fact, the oxidation state of the iron should be already III in 4a).

12

B. Meunier á J. Bernadou

2.2.5 Second Protonation Step (Generation of an Electrophilic Iron-Oxo Species) As for the ®rst protonation step, the proton which is necessary to produce an iron(V)-oxo species and a water molecule after protonation of the terminal oxygen atom of the FeAOOH entity is also provided by Thr-252, also involving the contribution of Asp-251, Lys-178, and Arg-186. This protonation is probably assisted by the negative charge accumulated on the proximal cysteine in the ®rst reduction step (see Sect. 2.2.1) which is one of the driving forces of the heterolytic cleavage of the OAO bond which generates the electrophilic high-valent FeV-oxo species. Since the protonation steps are faster than the second reduction step in wild P450s, no life-times can be provided for the FeAOOH and Fe@O species. Both are very short-life entities. For this reason, the exact nature of the hydroxylating species in cytochrome P450 is still a matter of debate. However, recent data presented in meetings suggest that the high valent iron-oxo species might be fully characterized by crystallography (Sligar et al., 9th International Conference on Biological Inorganic Chemistry, Minneapolis, 1999). In this second protonation step, the formal oxidation state of the iron center increases from III to V with the formation of an iron-oxygen double bond (Scheme 7). The electronic structure of the iron(V)-oxo species can be described as a singlet state (structure F1 in Scheme 7) and is probably the ground state of the active electrophilic species in P450. However, recent elegant calculations by Shaik et al. propose that the reactive species is a triplet state with a single ironoxygen bond and a strong radical character on the oxygen atom (structure F2) [38]. The third possible structure F3 corresponds with a transfer of one electron from the porphyrin ring to the metal center. Such FeIV-oxo radicalcation species has been fully characterized in heme-peroxidases (the so-called Compound I, see Sect. 2.5 and [39] for a recent book on peroxidases). These three high-valent structures correspond to a perferryl, an iron(IV)-oxyl or a ferryl porphyrin radical-cation state for F1, F2, or F3, respectively. The iron(V)oxo species F1 has not been characterized at the enzyme level, but with synthetic metalloporphyrins, in fact only with manganese derivatives up to now (see Sect. 2.4 on manganese(V)-oxo porphyrin complexes).

Scheme 7. Representation of the different possible structures of the electrophilic high-valent

iron-oxo species in cytochrome P450

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

13

It should be noted that an iron(III)-oxene is another possible limit form for the representation of the active electrophilic species of P450. However, such an electronic structure has not been observed when oxidizing iron(III) porphyrins with oxygen atom donors (in all cases the oxidation state of the iron center is at least IV as determined by magnetic data on isolated complexes, see [6±10]). Consequently, there is no longer any reason to use the name ``iron-oxene'' for the reactive iron-oxo species of P450 enzymes. The possible formation of an iron-oxo species as the key intermediate in oxygenation reactions catalyzed by cytochrome P450 has been evidenced by using a single oxygen atom donor as oxygen source with the enzyme itself [3] or with synthetic metalloporphyrins used as P450 chemical models [6]. We will discuss now the main aspects of the mechanism of the hydroxylation of aliphatic CAH bond by P450. Two main mechanisms have been characterized depending on the nature of the substrate and on the enzyme itself: a cagecontrolled radical mechanism (the ``oxygen rebound'' mechanism, see [6]) or a concerted mechanism without formation of a radical intermediate based on the use of ultrafast ``radical clocks'' (see [40, 41] for recent works on this aspect). Both possible reaction pathways are depicted in Scheme 8. Hydroxylation reactions catalyzed by P450s are not always stereospeci®c; the loss of stereochemistry has been shown with a tetradeuterated norbornane derivative and with many other substrates [2, 42]. The partial loss of retention of con®guration can easily be explained by the oxygen rebound mechanism with a rate of the formation of the carbon-oxygen bond being very close to the

Scheme 8. Two possible reaction pathways in hydroxylation reactions catalyzed by P450

enzymes: a concerted mechanism and a cage controlled radical pathway (the ``oxygen rebound'' mechanism)

14

B. Meunier á J. Bernadou

rate of the inversion of con®guration of the intermediate radical generated after abstraction of an H-atom by the high-valent iron-oxo species (see Scheme 9). The abstraction of a hydrogen atom by the active species of P450 is consistent with the high intrinsic isotope effects observed in hydroxylation reactions catalyzed by these heme-monooxygenases [2, 43, 44]. The kH/kD values range from 5 to 12, depending on the substrates and the category of P450. In addition, it should be noted that the value of the intrinsic isotope effect of the hydroxylation step can be masked by the other enzymatic steps, reversible substrate binding, product release, etc. In fact, the KIE values determined with P450 and t-BuO (a pure H-atom abstractor) in a series of different substrates are very similar, suggesting that both reagents have in common an H-atom transfer step [44]. Kinetic data on the hydroxylation reaction catalyzed by P450s have been obtained by using several different radical clocks to probe the rate of the oxygen atom transfer step [40, 41]. The absence of ring opening products in the hydroxylation of trans-1-methyl-2-phenylcyclopropane (the rate constant for ring opening is 2 ´ 109 s)1) with two different P450s (from rat or rabbit) indicates that the rate constant for the formation of the CAO bond in the alcohol product should be above 2 ´ 1011 s)1 [40]. Using related cyclopropane derivatives, Newcomb et al. estimated that the maximum lifetime of a putative C-centered radical intermediate in the hydroxylation by P4502B1 should be less than 1 picosecond [41]. These kinetic data strongly indicate that substrateprotein interactions within the active site (the enzyme is ``holding'' the substrate) probably have a key role in reducing the ring-opening rates of these radical clocks compared to their rates in solution and also contribute to enhance the rates of the CAO bond formation above rate values controlled by diffusion in solution. The role of an iron(V)-oxo species as being the only species responsible for electrophilic oxygen atom transfers (e.g., hydroxylation and epoxidation) has been questioned recently by Vaz and Coon based on studies with P450 mutants in which the threonine involved in the proton delivery within the active site was replaced by alanine. They obtained two contradictory results. Epoxidation and hydroxylation rates with the T302 A mutant of P4502B4 were

Scheme 9. Possible loss of retention of con®guration in hydroxylation reactions involving an

oxygen rebound mechanism

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

15

reduced whereas epoxidation rates were increased with the T303 A mutant of P4502E1 [45, 46]. These data have been interpreted as evidence of the electrophilic properties of the FeIIIAOOH intermediate (this species being reported as being able to behave as nucleophile in aldehyde deformylation or as electrophile in epoxidation). However, the observed differences in product formation between the wild and the mutant enzyme can be explained by an enhanced proton delivery by bulk water as observed in the Asp-251 mutant of P450cam [29] and also by a facilitated approach of the ole®n to the iron-oxo species. The mechanistic proposals by Newcomb, Coon, and Vaz are summarized in Scheme 10. The fact that the iron(III)-hydroperoxo intermediate still has a negative charge in its coordination sphere makes it a poor candidate as electrophilic agent. The distribution of charges on the P450 intermediates described in Scheme 10 is rather confusing in terms of the formal classical rules used in coordination chemistry for the description of metal complexes. The presence of two negative charges on the terminal oxygen of the iron-peroxo species is not correct: this terminal oxygen atom being already engaged in a covalent bond with the other oxygen atom, it can carry only one negative charge. In fact, the second negative charge is delocalized on the proximal cysteinato ligand (see Sect. 2.2.1) and should be indicated outside of the complex as described in Scheme 11. The iron(III)-peroxo species with two negative charges is probably the ``super-nucleophile'' in P450 intermediates (see above for the mechanism of aldehyde deformylation). Then the ®rst proton, delivered by the threonine or directly by bulk water, is added to the terminal negatively charged oxygen atom of the peroxo which is a more basic site than the proximal sulfur with its delocalized negative charge. The iron(III)-hydroperoxo species obtained after this ®rst protonation step still has one negative charge which is on the side of the proximal cysteine and not on the oxygen atom of the hydroperoxo ligand (see the structure of the nucleophilic iron(III)-hydroperoxo in Scheme 11).

Scheme 10. Different oxidant species generated by P450 enzymes according to Vaz et al.

Reproduced with permission from [46]

16

B. Meunier á J. Bernadou

Scheme 11. Formal descriptions of iron(III)-peroxo, iron(III)-hydroperoxo, and iron(V)-oxo

species with indication of the negative charges

The second proton delivery by the threonine to the hydroperoxo ligand gives two neutral molecules, the iron(V)-oxo and a water molecule. The development of DFT calculations, or related methods, will certainly be very helpful to describe in terms of molecular orbitals the electronic structures and the properties of these different reactive P450 intermediates which cannot be easily isolated and characterized by classical spectroscopic methods [47, 48]. We will now report a few recent examples of the reactivity of metalhydroperoxo and metal-oxo porphyrin complexes. This non-exhaustive presentation will only provide the necessary background related to reactive intermediates of the catalytic cycle of cytochrome P450 enzymes. 2.3 Reactivity of Iron(III)-Peroxo Porphyrin Complexes

The nucleophilicity of iron(III)-peroxo porphyrin complexes has been recently illustrated by Valentine et al. [49, 50]. The high reactivity of negatively charged iron(III)-peroxo species having a side-on structure has been demonstrated by using electron de®cient ole®ns (unsaturated ketones or quinones). The peroxo complex [FeIII(TMP)O2]) reacts quickly with menadione (2-methyl-1,4-naphthoquinone) to produce the corresponding 2,3-epoxide whereas no reaction was observed with cyclohexene [49]. In contrast, this latter electron-rich ole®n is easily epoxidized by single oxygen atom donors (iodosylbenzene, hypochlorite, monopersulfate) using metalloporphyrin as catalysts [6, 7]. When replacing the electron-donating porphyrin ligand by an electronwithdrawing porphyrin containing per¯uorophenyl substituents at the meso positions, the nucleophilicity of the corresponding iron-peroxo is reduced. But the presence of an axial ligand can restore its reactivity by opening the side-on peroxo to a more reactive end-on structure [50]. The peroxo complex [FeIII(TMP)O2]) is also able to mimic the deformylation step of aromatase [51]. What do we need to modify the reactivity of an iron-peroxo porphyrin complex, to go from a nucleophile to an electrophile active species able to epoxidize electron-rich ole®ns? In order to remove electron density from the

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

17

iron-peroxo, highly electron-de®cient porphyrin ligands should be used [52] associated with an electron-withdrawing proximal ligand. These properties are both the opposite of the electronic properties of the prosthetic group of P450; the natural porphyrin PPIX is a very good electron reservoir and the proximal cysteine ligand, associated to a network of amino acids, is also able to accumulate electron density. Both factors are strongly against (in favor of) electrophilic (nucleophilic) behavior of iron-peroxo intermediates in cytochrome P450s. It should be noted that non-heme metal-peroxo complexes which are able to react with electron-rich ole®ns (di- or tri-substituted aliphatic ole®ns) have electron-attracting ligands in their coordination sphere. This is the case for rhenium-oxo-peroxo catalysts (see [53] for a recent article, and the chapter by Herrmann and KuÈhn in this volume) or of iron-peroxo polypyridine complexes when the iron center is still able to behave as a Lewis acid with respect to the terminal oxygen atom of the peroxo motif (see [54], and the chapter by Girerd, Banse and Simaan in this volume). 2.4 Characterization and Reactivity of High-Valent Metal-Oxo Porphyrin Complexes

The ®rst high-valent iron-oxo porphyrin complex has been isolated by Groves et al. by oxidation at low temperature of Fe(TMP)Cl with m-chloroperbenzoic acid [55]. The resulting ``green'' compound is an iron(IV)-oxo complex with a radical-cation on the porphyrin ring which resembles the Compound I of heme-peroxidases (see [6, 7, 56] for review articles on the preparation and characterization of high-valent iron-oxo porphyrin complexes). Many other Compound I models have been obtained with various differently substituted metalloporphyrins. However, in all these model systems, the exchange between the spin S = 1 of the ferryl group with the spin S¢ = 1/2 of the porphyrin radical-cation was found to be strongly ferromagnetic in contrast to the weak ferromagnetic coupling usually observed with Cpd I derivatives [57]. Up to now, nobody has been able to identify an iron(V)-oxo. One possible explanation is probably related to the fact that most of these high-valent ironoxo-complexes have no proximal anionic ligand or a neutral one as in horseradish peroxidase which has a proximal histidine [58]. The P450 models with a proximal sulfur-containing ligand are more reactive than the corresponding models without this cysteine analogue and they might be the only suitable P450 models to allow the isolation of a true perferryl complex [59]. Oxygen-containing proximal ligands reduce the oxygenase activity of synthetic metalloporphyrins compared to the corresponding complexes with nitrogen-containing axial ligands [60]. The electron reservoir capacity of porphyrin ligands is in fact highly adapted to favor the formation of high-valent oxidation states of a metal center, still keeping a suf®cient lability of the metal-oxygen bond in order to obtain a facile oxygen atom transfer. The stability of high-valent iron- or manganese-oxo species is highly dependent on the nature of the ligands. For example, an inert d2 square-pyramidal manganese(V)-oxo stable at room temperature has been

18

B. Meunier á J. Bernadou

prepared with a diamido ligand [61, 62]. An additional example of lability vs stability of high-valent species is the case of nitrido-manganese(V) porphyrin complexes which are kinetically inert with organic substrates, but are able to transfer the nitrido motif to chromium(III) porphyrin via a two-electron redox process mediated by a heterobimetallic l-nitrido intermediate [63]. The same nitrido-manganese(V) porphyrin complexes can be transformed into nitrogen atom-transfer agents after acylation of the nitrido ligand to generate a labile acylimido-manganese(V) porphyrin complex [64]. Both MnIV-oxo and MnV-oxo porphyrin complexes have recently been characterized (see [65, 66] for early attempts). A (Por)MnIV@O complex has been characterized by X-ray absorption spectroscopy [67]. The MnAO bond Ê and the complex has a S = 3/2 spin state corresponding to a distance is 1.69 A high spin d3 con®guration. More recently, it has been possible to characterize a manganese(V)-oxo porphyrin complex by stopped-¯ow spectrophotometry [68]. Produced with hypochlorite or monopersulfate, (TMPyP)(X)MnV@O exhibits a Soret band at 443 nm between that of (TMPyP)MnIV@O (428 nm) and MnIII(TMPyP) (462 nm). The half-lifetime of this MnV-oxo is 25 ms at room temperature at pH 7.4. Its decay to the MnIV-oxo species can be accelerated by nitrite. When the methyl groups of the pyridinium substituents of the porphyrin ligand are in ortho positions instead of para, then the manganese(V)-oxo porphyrin complex has a longer lifetime (a few minutes) making possible the acquisition of a diamagnetic proton NMR spectrum [69]. This fact supports a low-spin d2 electronic structure for the ground state of this manganese(V)-oxo complex. This ``singlet'' state is probably in equilibrium with a ``triplet'' state; i.e., an MnIV-oxyl species with a single oxygenmanganese bond, as proposed by Shaik et al. [38]. In fact, the electronic structure of a metal(V)-oxo complex with a metal-oxygen double bond is probably higher in energy than the ``diradical'' form, a metal(IV)-oxyl species, as singlet dioxygen, with a double bond, is higher in energy by 23 kcal mol)1 compared to its triplet ground state, with a single OAO bond and two unpaired electrons (the name ``oxyl'' underlines the reduction of the bond order between the metal center and the oxygen atom). For a schematic representation of these metal-oxo vs metal-oxyl species, see structure F1 and F2, respectively in Scheme 7. A spin conversion of the singlet to the triplet state of the metal(V)oxo species will occur during the oxygen atom transfer step [38]. The capacity of these electrophilic high-valent iron-oxo and manganese-oxo porphyrin complexes to transfer an oxygen atom to hydrocarbons or ole®ns will be described with recent mechanistic details in Sect. 3. The existence of MnV-oxo entities has now been evidenced both by spectroscopic methods and reactivity data. 2.5 Active Species in Heme-Peroxidase Catalytic Cycles

Several heme-peroxidases from the vegetable or animal kingdoms have been extensively studied: e.g., horseradish peroxidase, chloroperoxidase, ligninase, myeloperoxidase, lactoperoxidase, etc. (see [39] for a book and [70, 71] for

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

19

recent reviews). Here we will only discuss some recent aspects concerning the reactivity of the different possible intermediates (metal-oxo vs metal-peroxo, electron transfer vs oxygen atom transfer) generated during the catalytic cycles of these heme-peroxidases. We will limit the discussion to horseradish peroxidase HRP and chloroperoxidase CPO. 2.5.1 Horseradish Peroxidase HRP catalyzes the abstraction of one or two electrons (usually two electrons) via a single electron transfer from an organic substrate, hydrogen peroxide being used as electron acceptor, according to Eq. (2): peroxidase

AH2 ‡ H2 O2 ƒƒƒƒƒƒƒƒƒ! A ‡ 2H2 O

…2†

The different oxidation states and the properties of this enzyme have been well studied. The prosthetic group of HRP is a ferric-protoporphyrin IX with a histidine as proximal ligand (His-170). The high-valent iron-oxo entity of HRP-Compound I is generated from hydrogen peroxide via its heterolytic cleavage involving the guanidinium function of Arg-38 (acting as a proton source for the formation of the water molecule) and His-42 acting as a base in the deprotonation step of H2O2 (Scheme 12) [58, 72]. The oxidation of HRP by hydrogen peroxide generates a reactive intermediate (Compound I) containing two redox equivalents above the resting state of the native enzyme. The one-electron reduction of Compound I by the substrate generates Compound II which is still able to abstract one electron from the substrate. Both high-valent iron species have been characterized by several physico-chemical methods. Compounds I and II have a broad Soret band at 400 nm and 420 nm, respectively. Compound I consists of an iron(IV)-oxo species with a radical cation on the porphyrin ring (Scheme 13). Compounds I and II have been characterized by X-ray absorption and Raman spectroscopies. Both methods con®rmed the presence of an iron(IV)Ê oxo entity in both Compounds I and II with a short Fe@O bond of 1.6 A

Scheme 12. Essential amino acids involved in the activation of hydrogen peroxide by

horseradish peroxidase

20

B. Meunier á J. Bernadou

Scheme 13. Catalytic cycle of horseradish peroxidase (AH2 being a 2-electron donor)

consistent with a ferryl structure [73]. More recently, it has been suggested Ê at pH 7 [74]. that the FeAO bond distance of Compound II was longer: 1.9 A Resonance Raman studies con®rmed the presence of an Fe@O entity for both Compounds I and II with vibrations at 737 cm)1 and 776 cm)1, respectively [75]. The p-radical-cation of Compound I has a predominant 2A2u. This porphyrin radical is ferromagnetically coupled with the spin S = 1 of the ferryl state [76]. The lifetime of HRP-Compound I can be increased up to one hour at )20 °C with a polyethylene glycolated enzyme [77]. Compound III is an inactive intermediate corresponding to an iron(III)-peroxo species resulting from the addition of dioxygen to the ferrous state of HRP or from the reaction of an excess of hydrogen peroxide with the ferric state of the enzyme [78]. The high-valent iron-oxo intermediates of HRP are not directly accessible to the different substrates. Alkylation of the d-meso position of the heme group by alkylhydrazines indicates that substrates approach the active site of HRP from one edge of the prosthetic group [79]. Recent data on the substrate binding site of HRP have been obtained by proton NMR spectroscopy and molecular dynamics [80, 81]. The weak binding site of HRP is an open pocket able to accommodate a large range of substrates. Despite several claims, HRP is unable to catalyze the transfer of an oxygen atom from hydrogen peroxide to a substrate molecule. HRP is unable to oxidize styrene whereas chloroperoxidase is able to catalyze the oxidation of this ole®n to a mixture of the corresponding epoxide and phenylacetaldehyde [82, 83]. The oxidation of sul®des by HRP involves the formation of a radical-cation on the sulfur atom followed by water addition, not an oxygen atom transfer [84]. Despite the use of hydrogen peroxide, a suitable oxidant to generate metalhydroperoxo species, it should be noted that all the known intermediates involved in the HRP catalytic cycle are high-valent iron-oxo entities. 2.5.2 Chloroperoxidase The fungal chloroperoxidase (CPO) is among the few peroxidases which are able to catalyze the oxidative chlorination of substrates using H2O2 and Cl) (myeloperoxidase is an other example) according to Eq. (3) [85]:

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases CPO

AH ‡ H2 O2 ‡ Cl ‡ H‡ ƒƒƒƒ! ACl ‡ 2 H2 O

21 …3†

This oxidative chlorination is performed on substrates containing an activated carbon-hydrogen bond such as b-diketones (chlorodimedone is a classical substrate used in CPO assays). A noticeable difference between CPO and HRP is the presence of a cysteine residue as proximal ligand (Cys-29 in CPO) instead of a histidine in HRP [86]. The X-ray structure has been determined recently and con®rms the presence of a manganese(II) ion bound to a heme propionate and also surrounded by His-105, Glu-104, and Ser-108 (the manganese ion may be the binding site for a chloride ion) [87, 88]. The addition of hydrogen peroxide to the ferric state of the enzyme generates CPO-Compound I, the only detectable intermediate having an FeIV@O bond with a Raman stretching band at 790 cm)1 [89]. The radicalcation of Compound I is probably delocalized on the macrocycle and on the axial ligand. The addition of chloride to Compound I might generate an FeIII-OCl entity, a possible candidate to explain the substrate chlorination [90]. An alternative mechanism is the formation of free HOCl. The same mixture of chloroanisole isomers was observed in the chlorination of anisole by free HOCl or by the CPO chlorinating system [91]. A third possible mechanism is to consider that the manganese site of CPO is able to facilitate the binding of Cl). The short distance between this Mn-site and the heme should facilitate the electron removal from the chloride ion and then Cl+ could be transferred to the substrate present near the active site. This hypothesis would explain the absence of formation of free HOCl without invoking the formation of an iron(III)-hypochlorito intermediate during the catalytic cycle of chloroperoxidase. Chloroperoxidase is inactivated by formation of N-alkyl-heme in the oxidation of terminal ole®ns [92]. These data suggest that different substrates have relatively easy access to the active site of CPO. Unlike HRP, CPO is able to catalyze the epoxidation of different ole®ns (styrene, propylene, allyl chloride) [93]. The oxygen atom of the epoxide arises from the primary oxidant as evidenced by using 18O-labeled hydrogen peroxide [83]. Recent studies have demonstrated that the CPO-catalyzed epoxidation of cis-disubstituted ole®ns is enantioselective; enantiomeric excesses (ee values) range from 33% to 85% [93]. With a slow addition of hydrogen peroxide to avoid the degradation of CPO, it has been possible to obtain 4200 catalytic cycles with an ee value of 94% in the epoxidation of methylallyl propionate [94]. Chloroperoxidase is also able to catalyze the enantioselective oxidation of sul®des to provide chiral sulfoxides with ees up to 90±95% [95, 96]. It has been proposed from data obtained with a series of para-substituted thioanisoles that S-oxygenations catalyzed by CPO involve an oxygen atom transfer from Cpd I to the substrate rather than a one-electron oxidation of the sul®de followed by the addition of H2O on the S-radical-cation intermediate, as proposed for sul®de oxidations catalyzed by HRP [97]. Again, no iron-peroxo species has been identi®ed among the active species generated with hydrogen peroxide during the catalytic cycle of chloroperoxidase.

22

B. Meunier á J. Bernadou

Recent claims mentioned that microperoxidase might be able to oxygenate substrates via an iron-peroxo species generated by addition of an hydroxide to an iron-oxo (this formation of a weak OAO bond is certainly thermodynamically not favorable; in fact this is the reverse reaction of the heterolytic cleavage of the peroxidic bond when hydrogen peroxide is activated by peroxidases) [98]. However, these claims are based on labeling experiments obtained from a rather complicated work-up on few catalytic cycles. This work should be independently con®rmed in effective catalytic reactions before claiming that iron-peroxo species are involved in oxygenation reactions mediated by microperoxidase (see footnote 48 in [41]).

3 Oxo-Hydroxo Tautomerism with Water-Soluble Metalloporphyrins In the chemistry of P450 models, high valent metal-oxo species have been developed as active intermediates in many different oxidation reactions using manganese or iron porphyrin catalysts and oxygen atom donors (PhIO, NaOCl, KHSO5, H2O2, etc), or dioxygen associated to a reductant, as oxygen atom source. When the oxygenation reaction (hydroxylation or epoxidation) is performed in an organic solvent (usually dichloromethane) with hydrophobic metalloporphyrin catalysts, the oxygen atom incorporated within the substrate originates from the oxidant. This has been evidenced for hypochlorite [66, 99], monopersulfate [100], or iodosylbenzene [101 and references therein], but when these metalloporphyrin-catalyzed oxygenations are performed in aqueous solvents, such metal-oxo species are able to transfer an oxygen atom coming from either the oxygen source or from bulk water. Because the intermolecular exchange of metal-oxo with bulk water is slow, an intramolecular exchange of labeled oxygen atoms via the so-called oxo-hydroxo tautomerism [102] has been proposed [103]. This mechanism involves a rapid shift of two electrons and one proton from a hydroxo ligand (electron-rich ligand formed by deprotonation of an aqua ligand) to the trans oxo species (electron-poor ligand) leading to the transformation of the hydroxo ligand into an electrophilic oxo entity on the opposite side of the initial oxo (see Scheme 14 for a representation of this tautomeric equilibrium).

Scheme 14. Key step of the oxo-hydroxo tautomerism

The oxo-hydroxo tautomerism can contribute to a better characterization and an improved understanding of chemical reactivities of these high-valent metal-oxo species, not only important for the knowledge of heme-enzymes

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

23

catalyzing oxidations, but also in the design of ef®cient biomimetic or bioinspired oxidation catalysts. At the present stage of the knowledge, the design of metal-oxo complexes able to catalyse oxygenations ef®ciently is still challenging, since all the parameters involved in the different O-atom transfer steps are not fully understood. 3.1 A New Type of Tautomerism: The Oxo-Hydroxo Tautomerism

We recently discover the oxo-hydroxo tautomerism during studies on the origin of the epoxidic oxygen atom in an ole®n epoxidation performed in aqueous media with a water-soluble metalloporphyrin catalyst [103]. Using isotopically labeled oxidants or labeled water, it has been shown that the oxygen of the metal-oxo porphyrin complex can be quickly exchanged with water via the axial hydroxo ligand with reaction rates depending on the experimental conditions (pH, temperature, composition of the medium, nature of the axial ligands, etc.). We shall see that, using the label distribution in oxidation products modulated by this oxo-hydroxo tautomerism, it is possible to unambiguously distinguish between oxygenation reactions occurring via an O-transfer from a high-valent metal-oxo complex or via an autoxidation mechanism (see [104] for a review on this controversial debate) in metalloporphyrin-catalyzed oxygenations carried out in the presence of H2 18 O. The main consequence of the oxo-hydroxo tautomerism is the incorporation within the substrate of an oxygen atom coming from either the oxidant or from bulk water, respectively in the ratio 1:1 (Scheme 15). The degree of 18O-incorporation observed into the product (epoxide, alcohol, etc.) is then mechanistically informative.

Scheme 15. The oxo-hydroxo tautomerism mediates the incorporation into the oxidation

product of 50% of oxygen coming from the primary oxidant and 50% from water. X = hydroxo ligand

3.2 Some Previous Data on O-Exchange of Metal-Oxo Species with Bulk Water

Over the last 15 years, few reports have mentioned the use of 18O-labeling experiments in order to characterize high-valent metal-oxo species or to elucidate the mechanism of O-transfer by these reactive species. For reactions

24

B. Meunier á J. Bernadou

performed in the presence of water, results range from 0% to 100% of Oincorporation from water, depending on the experimental conditions. Most of these results can be understood by reference to the oxo-hydroxo tautomerism (for detailed comments, see the paragraph on requested conditions to observe oxo-hydroxo tautomerism) or on the basis of a direct, intermolecular Oexchange of the oxo ligand with the bulk water. It should be noted that the rate of this intermolecular exchange is slower than the rate of the oxo-hydroxo tautomerism which has been estimated to have a rate constant of 103 s)1 [68]. A resonance Raman investigation on Cpd II of HRP at pH 7, based on isotopic shift of the FeIV@O stretching mode of Cpd II (the mFe@O was observed at 774 cm)1 after activation in H2 16 O with either H2 16 O2 or H2 18 O2 , at 740 cm)1 for activation in H2 18 O with either H2 16 O2 or H2 18 O2 ) provided evidence for the oxygen atom exchange between the heme-FeIV@O and bulk water [105]. In a controversial manner, this exchange was not observed in the case of stable species of diacetylheme or manganese substituted HRP: the Raman spectrum of Cpd II generated with H2 16 O2 or H2 18 O2 presented lines at 781 cm)1 or 745 cm)1 (diacetylheme HRP) and at 626 cm)1 or 596 cm)1 (MnHRP), respectively. No isotope-induced change was observed at neutral pH for 1 h at 4 °C in the presence of H2 16 O or H2 18 O, indicating no appreciable exchange of the iron-oxo entity with bulk water [106]. In studies performed at )80 °C on a ferryl porphyrin p-cation-radical derived from the synthetic FeIII(TMP)Cl, the mFe@O band was observed at 828 cm)1 (activation with 16O-m-CPBA) or 792 cm)1 (activation with 18O-mCPBA), the ®rst band remaining unshifted in the presence of H2 18 O indicating that the oxygen atom of the iron-oxo was not easily exchanged with water under these conditions [107]. Incorporation of 18O from bulk water was initially reported for epoxidation of norbornene with m-CPBA catalyzed by FeIII(TMP)Cl in an organic medium containing 1% H2 18 O [55] and was further noticed in the course of epoxidation of b-methylstyrene by manganese(V)-oxo porphyrin in CH2Cl2 saturated in H2 18 O [108], indicating a rather fast exchange of the oxo ligand with water. This 18O-exchange was clearly slower in the case of manganese(IV) species and was inhibited by the presence of pyridine as axial ligand [109]. 3.3 Oxo-Hydroxo Tautomerism Observed in Different Metalloporphyrin-Catalyzed Oxygenation Reactions

3.3.1 Epoxidation Reactions We initially reported the oxo-hydroxo tautomerism to explain isotopic results observed in aqueous phase during KHSO5 epoxidation of carbamazepine (CBZ), an analgesic and anticonvulsant drug, catalyzed by a cationic watersoluble manganese porphyrin [103]. In such reaction performed at pH 5 in aqueous solution with various contents of H2 18 O, it was shown that half of the oxygen atoms incorporated in the epoxide came from the solvent (Fig. 1a). It

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

25

Fig. 1. a The amount of labeled oxygen found in CBZ oxide correlates with half the content

of 18O-label of water present in the reaction mixture. In abscisse: content (%) of H218O in water (reproduced with permission from [103]; see also [111] for a similar correlation). b Dependence of the oxo-hydroxo tautomerism on a suf®cient concentration of water in a non-aqueous solvent observed in cyclooctene catalyzed epoxidation. In abscisse: 0.5, 1, 1.5, 2, and 2.5 mmol of H218O correspond to 1, 2, 3, 4, and 5 molar concentrations of H218O, respectively. Substrate concentration was 20 mmol l)1 (reproduced with permission from [111])

was checked that neither CBZ-oxide nor KHSO5 [100, 110] exchanged oxygen atoms with water in the reaction conditions. To explain the ratio of 0.5 for the incorporation of oxygen from the solvent, we proposed an oxo-hydroxo tautomerism (previously named ``redox tautomerism'' in [103, 110±114]) involving a coordinated water molecule on the manganese(III) porphyrin precursor 1 (Schemes 14 and 16). It must be noted that a constant 50% O-exchange corresponds only to the oxo-hydroxo tautomerism involving the hydroxo ligand trans to the high-valent metal-oxo

Scheme 16. General scheme for the oxo-hydroxo tautomerism

26

B. Meunier á J. Bernadou

complex, but not to a direct exchange of the oxo ligand with bulk water that can lead to 100% O-exchange. MnIII(TMPyP), the pentaacetate of the diaqua-manganese(III) derivative of meso-tetrakis(1-methyl-pyridinium-4-yl)porphyrin, can exist in aqueous medium with one or two metal-bound water molecules as axial ligands [1 in Scheme 16; see [115] for an X-ray structure of the bis-aqua-Mn(TMPyP) complex]. Increasing the metal oxidation state from III to V (going from the MnIII complex 1 to the MnV-oxo 2) should lower suf®ciently the pKa value of the ligated water to allow, at the pH of the reaction, its conversion into a hydroxo ligand (3; see [116] for a discussion on the pKa values of aqua and hydroxo ligands in high-valent metalloporphyrins). Removal of a proton from this hydroxo ligand results in the formation of the stabilized anion 4 with 4e) delocalized on both metal-oxygen bonds (4¢ is a mesomeric form with 3e) delocalized and manganese at the formal oxidation state IV). This anion can be protonated with the same probability at the end of one of the two metal-oxolike bonds, giving rise to either form 3 or 5, which reacts with CBZ to produce CBZ-oxide containing either 16O or 18O, respectively, in the ratio 1 to 1. The conversion of 3 to 5 does not necessarily involve 4 and 4¢ as discrete deprotonated intermediates but might also proceed via a hydrogen-bonded water molecule in a more concerted fashion (Scheme 14). Other recent reports support the concept of oxo-hydroxo tautomerism [68, 111, 114]. Groves et al. [68] also reported the 18O-incorporation in CBZ-oxide in an Mn(TMPyP)-catalyzed oxidation of CBZ via an oxo-hydroxo interconversion. Lee and Nam described similar 18O-incorporation in cyclooctene epoxidation by either m-CPBA, H2O2 or t-BuOOH catalyzed by mesotetrakis(penta¯uoro-phenyl)porphyrinato-iron(III) chloride FeIII(F20TPP)Cl, the reaction being performed in a CH3OH/CH2Cl2 mixture containing 10% H2O [111]. Even at low pH values, CBZ epoxidation data obtained in aqueous solutions by H2O2, t-BuOOH, or KHSO5 in the presence of FeIII(TDCPPS) indicate that an oxo-hydroxo tautomerism was involved [114]. These latter experiments strongly supported that a common high-valent iron-oxo species was generated from the different oxidants and was the active species responsible for ole®n epoxidation. 3.3.2 Hydroxylation Reactions The oxo-hydroxo tautomerism was further characterized in the oxidation of deoxyribose CAH bonds of DNA by the Mn(TMPyP)/KHSO5 system (see Scheme 17) [112]. Hydroxylation at carbon-1¢ of deoxyribose gave in several steps 5-methylene-2-furanone (5-MF) as ®nal sugar residue. In the presence of labeled H2 18 O, 50% of oxygen coming from the primary oxidant (16O from KHSO5) and 50% from the solvent (18O from H2 18 O) were incorporated in 5MF, strongly supporting a metal-oxo mediated DNA cleavage with an oxohydroxo tautomerism to explain the 18O-incorporation in the desoxyribose oxidation product.

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

27

Scheme 17. Examples of oxo-hydroxo tautomerism from literature data. O: unlabeled

oxygen; d: labeled oxygen; f: mixed labeled oxygen

Another example came from the monopersulfate oxidation of 4-isopropylbenzoic acid performed in H2 18 O and catalyzed by the same water-soluble metalloporphyrin Mn(TMPyP) [110]. In the primary hydroxylation product, 4-(1-hydroxy-1-methylethyl)benzoic acid, nearly half of the oxygen atoms incorporated in the alcohol function came from water. In the cyclohexane hydroxylation by m-CPBA catalyzed by Fe(F20TPP)Cl [111], or by H2O2 catalyzed by [Fe(TF4TMAP)](CF3SO3)5 [117], the percentage of 18O incorporated in cyclohexanol was also found to be 40±50%, with a reaction mixture containing only 7±10% of H218 O. Oxo-hydroxo tautomerism was recently observed in both epoxidation and hydroxylation reactions catalyzed by metalloporphyrins using activation by the sul®te/dioxygen system. Oxidation of sul®te catalyzed by the water-soluble Mn(TMPyP) allows generation of MnV@O complex [see Scheme 18(A)]. This alternative, biocompatible oxidation system compared to the preformed oxidant KHSO5 allows incorporation of one labeled oxygen atom coming from either water or dioxygen [Scheme 18(B)] as was illustrated in the metalloporphyrin-catalyzed hydroxylation of a benzylic CAH bond [Scheme 18(C)] or in epoxidation of carbamazepine [118]. 3.3.3 Quinone Formation In an aqueous solution, the metalloporphyrin-catalyzed oxidation of 2-methylnaphthalene to p-quinones involves two consecutive oxygen transfers from an intermediate metal-oxo entity responsible for 30±55% indirect incorporation of 18O from water into the generated quinones (Scheme 17) [113].

28

B. Meunier á J. Bernadou

Scheme 18A±C. Activation of metalloporphyrins with O2 + SO23 : A mechanism of formation

of high-valent metal-oxo species; B the oxo-hydroxo tautomerism allows one to incorporate one labeled oxygen atom from either water or dioxygen. X = hydroxo ligand; C illustration of both labeling routes described in B when performing a metalloporphyrin-catalyzed hydroxylation of a benzylic CAH bond [118]. s: unlabeled oxygen; d: labeled oxygen; f: mixed labeled oxygen

3.4 Required Conditions to Observe an Oxo-Hydroxo Tautomerism

3.4.1 An Axial Ligand Competitive to the Hydroxo Ligand Inhibits Oxo-Hydroxo Tautomerism In heme-enzymes, the presence of a cysteinato ligand (cytochrome P-450 or chloroperoxidase) or an imidazole from an histidine (peroxidase) prevents water from coordinating to the iron center. This implies that there is no possible intramolecular exchange between high valent metal-oxo intermediates and water through an oxo-hydroxo tautomerism.

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

29

With synthetic metalloporphyrins the situation is different. Iron and manganese porphyrins are known to form imidazole and pyridine complexes when these heterocycles are present in reaction mixtures. In the epoxidation of cis-b-methylstyrene with m-CPBA catalyzed by MnIII(TMP)Cl, the presence of pyridine completely prevents isotopic enrichment of epoxide products when the reaction is performed in the presence of H2 18 O [108]. In the epoxidation of cyclooctene by Fe(F20TPP)Cl and H2O2, it was found that effectively the 18Oincorporation into the product diminished as the amount of 5-chloro-1methylimidazole added to the reaction mixture increased [111]. So when the axial position opposite to the oxo group is blocked by a ligand (not derived from water), the oxo-hydroxo tautomerism cannot occur and consequently the oxygen incorporated in the oxidation product is 100% from the oxidant, instead of 50% from the oxidant and 50% from water. In other respects, when the trans ligand is a water molecule instead of a hydroxo ligand, as is probably the case for metal(IV)-oxo species (see hereafter for a discussion on this point), the tautomerism is largely reduced. This may explain why the oxo-hydroxo tautomerism was not observed during the oxidation of polycyclic aromatic hydrocarbons catalyzed by iron tetrasulfophthalocyanine performed in the presence of H2 18 O (the reactive species was shown to be FeIV@O) [119]. 3.4.2 A Competitive Autoxidation Route Lowers Incorporation of Oxygen from Water Groves and Stern have noticed that during the cyclooctene [108] or cis-bmethylstyrene [109] epoxidation by manganese(IV)-oxo porphyrin under aerobic conditions, dioxygen was intimately involved in the oxidation process with oxidation products partially resulting from an autoxidation process. The consequence is a decrease of isotopic enrichment from solvent. During the monopersulfate oxidation of ketoprofen catalyzed by Mn(TMPyP), trapping of the intermediate C-centered radical on the substrate by molecular oxygen was competing with the oxygen rebound mechanism, explaining the observed reduction of 18O-incorporation from solvent in the ®nal product when ketoprofen was oxidized under air by the system Mn(TMPyP)/KHSO5/H2 18 O system [110]. 3.4.3 Temperature Effect In order to characterize the high-valent reactive species formed by oxidative activation of metalloporphyrins, many experiments have been performed at low temperature with some of them concerning 18O-incorporation into the products in the presence of H2 18 O. In the more extensive study, Lee and Nam [111] described the cyclooctene epoxidation by H2O2 or m-CPBA in the presence of Fe(F20TPP)Cl at different temperatures. The 18O-enrichment in the epoxide gradually increased as the reaction temperature raised from )78 °C to

30

B. Meunier á J. Bernadou

45 °C. The authors suggested that a putative FeIIIAOOR species was involved at low temperature, with a rate of the OAO bond cleavage (leading to the highvalent iron-oxo porphyrin complex) lower than the oxygen atom transfer rate (k1 < k3 in Scheme 19). At higher temperature the formation of the high-valent iron oxo porphyrin should be favored (increase of k1). However, since a tautomeric equilibrium is highly dependent on temperature, the present results might alternatively be reinterpreted as a fast formation of the iron-oxo species (k1 > k3), even at low temperature, but with a slow prototropy (k2 < k4) at this temperature and a faster one (k2 > k4) at higher temperature (k3 and k4 = oxygen atom transfer rates; the best conditions to observe the oxo-hydroxo tautomerism correspond to k1 > k3 and k2 > k4). We must note that even at low temperature ()78 °C) iron-oxo porphyrin species have been detected and characterized [55, 120, 121]. 3.4.4 Differences in Exchange Kinetics for Metal(IV)-Oxo and Metal(V)-Oxo From experiments conducted on cis-b-methylstyrene with manganese(V)-oxo and manganese(IV)-oxo porphyrin complexes, Groves and Stern concluded from the 18O results that, in addition to differences in oxygen transfer occurring with retention or loss of the stereochemistry, the manganese(IV)oxo slowly exchanged its oxo ligand with H2 18 O, while the exchange was very fast for the manganese(V)-oxo complex [108]. These data are consistent with Schemes 16 and 20 and with literature data on the proton acidity of coordinated water in high-valent species [116, 122, 123]. Going from 3 to 5 (Scheme 16) only requires a prototropy in the case of a manganese(V)-oxo species (Scheme 20; the oxo-hydroxo being the major form), whereas in the

Scheme 19. Kinetic parameters depending on temperature. X = hydroxo ligand

Scheme 20. Equilibria between oxo-aqua and oxo-hydroxo forms depending on the oxidation

state of the metal center

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

31

case of manganese(IV)-oxo the ligand trans to the oxo is mainly a water molecule due to the lower acidity of the ligated water when the metal oxidation state is reduced. So, for manganese(IV)-oxo complexes, the oxo-hydroxo tautomerism can only affect the small fraction of metal-oxo species with a hydroxo ligand, the oxo-aqua form being dominant in this case (Scheme 20). 3.4.5 Role of the Ratio Water/Substrate Concentrations The percentage of 18O incorporated in oxidation products might be governed by the relative rate to reach the tautomerism equilibrium (k2 in Scheme 19) which is function of the water concentration, and the rate of oxygen transfer (k4) which depends on the substrate concentration. For reaction performed in an essentially aqueous medium (a 90% aqueous solution is 50 mol l)1 in water) and a substrate rather diluted (1 mmol l)1 for example), the competition between the tautomerism and the oxygen transfer was not observed: the water/substrate molar ratio (5 ´ 104) was largely in favor of the oxo-hydroxo tautomerism [103]. In an organic medium containing only a small amount of water, and at high substrate concentrations, the situation is the opposite and can affect the tautomerism equilibrium and consequently the level of 18Oincorporation [114]. An example of such an extreme condition concerns the epoxidation of 1 mol l)1 cyclohexene solution in CH2Cl2/CH3OH containing 5% of H2 18 O with a metalloporphyrin catalyst and H2O2, t-BuOOH, or mCPBA as oxygen donor. In these conditions, no 18O-incorporation was observed in the epoxide (water/substrate molar ratio 3) [101]. In intermediate conditions, such as in epoxidation of 20 mmol l)1 cyclooctene with H2O2 catalyzed by Fe(F20TPP)Cl in an organic medium containing increasing amounts of water, the 50% O-incorporation from water was observed for a reaction mixture containing at least 10% of water (water/substrate molar ratio above 250; Fig. 1b) [111]. In the rare case of a very small concentration of substrate, then the reaction becomes very slow and the aqua ligand (form 2 or 6 in Scheme 16) can exchange with water from solvent and the 18O incorporation from solvent can rise above 50% [111]. 3.4.6 Other Parameters (pH, etc.) Several other parameters may probably in¯uence the oxo-hydroxo tautomerism but no systematic study has been done up to now. Among them, the pH value of the reaction mixture surely plays a key role in this prototropy mechanism as also do the nature of the metal and the ligand in the metalloporphyrins through their capacity to form differently coordinated complexes. In addition, some reported variations of the 18O rate of incorporation around 50% suggest that the oxo-hydroxo tautomerism equilibrium is probably tuned by small variations of kinetic parameters, including solvent effects or differences in the kinetic parameters of the oxygen transfer from the metal-oxo intermediate to the organic substrate.

32

B. Meunier á J. Bernadou

4 Conclusion More than 40 years after the discovery of cytochrome P450 enzymes, the exact nature of the active species (the putative high-valent iron-oxo intermediate) is still a matter of intensive debates. These scienti®c debates are in fact highly fruitful: enzymologists and structural biologists had to talk with chemists and vice versa, and data obtained with chemical models of these monooxygenases had to be compared with data obtained with the enzymes. Site-directed mutagenesis studies in relation with structural studies allowed one to check different hypotheses. All these studies have enriched the knowledge in molecular enzymology and the coordination chemistry of metal-peroxo and high-valent metal-oxo species. Among the different methods for studying the mechanism of oxidation mediated by high-valent metal-oxo complexes, the recent discovery of the ``oxo-hydroxo tautomerism'' provides an additional useful tool to discuss the mechanism of catalytic O-atom transfer reactions, the nature of the axial ligand trans to the oxo species, and the oxidation state of these metal-oxo entities. Acknowledgements. The authors are deeply indebted to the work of collaborators and co-

workers whose names are listed in several references of this chapter.

5 References 1. Nelson DR, Koymans L, Kamataki T, Stegeman JJ, Feyereisen R, Waxman DJ, Waterman MR, Gotoh O, Coon MJ, Estabrook RW, Gunsalus IC, Nebert DW (1996) Pharmacogenetics 6: 1 2. Ortiz de Montellano PR (ed) (1996) Cytochrome P450: structure, mechanism and biochemistry. Plenum, New York 3. Ullrich V (1979) Topics in Current Chemistry 83: 67 4. Taube H (1965) J Gen Physiol 49: 29 5. Hayaishi O (ed) (1974) Molecular mechanism of oxygen activation. Academic Press, New York 6. Groves JT, Han YZ (1996) In: Ortiz de Montellano PR (ed) Cytochrome P450: structure, mechanism and biochemistry. Plenum, New York, chap 1, pp 3±48 7. Meunier B (1992) Chem Rev 92: 1411 8. Mansuy D (1987) Pure Appl Chem 59: 759 9. Ostovic D, Bruice TC (1992) Acc Chem Res 25: 314 10. Dolphin D, Traylor TG, Xie LY (1997) Acc Chem Res 30: 259 11. Vaz ADN, McGinnity DF, Coon MJ (1998) Proc Natl Acad Sci USA 95: 3555 12. Meunier B (2000) Biomimetic oxidations mediated by metal complexes. Imperial College Press, London 13. Bernadou J, Meunier B (1998) Chem Commun 2167 14. Poulos TL, Finzel BC, Howard AJ (1986) Biochemistry 25: 5314 15. Poulos TL, Finzel BC, Howard AJ (1987) J Mol Biol 195: 687 16. Grif®n BW, Peterson JA (1972) Biochemistry 11: 4740 17. Atkins WA, Sligar SG (1990) Biochemistry 29: 1271

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

33

18. Poulos TL, Cupp-Vickery J, Li H (1996) In: Ortiz de Montellano PR (ed) (1996) Cytochrome P450: structure, mechanism and biochemistry. Plenum, New York, chap 4, pp 125±150 19. Ravichandran KG, Boddupalli SS, Hasemann CA, Peterson JA, Deisenhofer J (1993) Science 261: 731 20. Gleria KD, Nickerson DP, Hill HAO, Wong LL, FuÈloÈp V (1998) J Am Chem Soc 120: 46 21. Raag R, Poulos TL (1989) Biochemistry 28: 7586 22. Hanson LK, Eaton WA, Sligar SG, Gunsalus IC, Gouterman M, Connell CR (1976) J Am Chem Soc 98: 2672 23. Caron C, Mitschler A, Rivere G, Ricard L, Schappacher M, Weiss R (1979) J Am Chem Soc 101: 7401 24. Ueno T, Kuosumi Y, Yoshizawa-Kumagaye K, Nakajima K, Ueyama N, Okamura TA, Nakamura A (1998) J Am Chem Soc 120: 12,264 25. Egawa T, Ogura T, Makino R, Ishimura Y, Kitagawa T (1991) J Biol Chem 266: 10,246 26. MacDonald IDG, Sligar SG, Christian JF, Unno M, Champion PM (1999) J Am Chem Soc 121: 376 27. El-Kasmi D, Tetreau C, Lavalette D, Momenteau M (1995) J Am Chem Soc 117: 6041 28. Benson DE, Suslick KS, Sligar SG (1997) Biochemistry 36: 5104 29. Vidakovic M, Sligar SG, Li H, Poulos TL (1998) Biochemistry 37: 9211 30. Vaz ADN, Roberts ES, Coon MJ (1991) J Am Chem Soc 113: 5886 31. Roberts ES, Vaz ADN, Coon MJ (1991) Proc Natl Acad Sci USA 88: 8963 32. Caspi E, Arunachalam T, Nelson PA (1986) J Am Chem Soc 108: 1847 33. (a) Ortiz de Montellano PR (1996) In: Ortiz de Montellano PR (ed) (1996) Cytochrome P450: structure, mechanism and biochemistry. Plenum, New York, chap 8, pp 245±303; (b) Graham-Lorenze S, Amarneh B, White RE, Peterson JA, Simpson ER (1995) Protein Science 4: 1065 34. Moali C, Boucher JL, Sari MA, Stueher DJ, Mansuy D (1998) Biochemistry 37: 10,453 35. Ortiz de Montellano PR (1998) Acc Chem Res 31: 543 36. Gerber NC, Sligar SG (1992) J Am Chem Soc 114: 8742 37. Raag R, Martinis SA, Sligar SG, Poulos TL (1991) Biochemistry 30: 11,420 38. Shaik S, Filatov M, SchroÈder D, Schwarz H (1998) Chem Eur J 4: 193 39. Everse J, Everse KE, Grisham MB (eds) (1991) Peroxidases in chemistry and biology (2 volumes). CRC Press, Boca Raton 40. Atkinson JF, Ingold KU (1993) Biochemistry 32: 9209 41. Toy PH, Newcomb M, Hollenberg PF (1998) J Am Chem Soc 120: 7719 42. Groves JT, McClusky GA, White RE, Coon MJ (1978) Biochem Biophys Res Commun 81: 154 43. Iyer KR, Jones JP, Darbyshire JF, Trager WF (1997) Biochemistry 36: 7136 44. Manchester JI, Dinnocenzo JP, Higgins LA, Jones JP (1997) J Am Chem Soc 119: 5069 45. Vaz ADN, McGinnity DF, Coon MJ (1998) Proc Natl Acad Sci USA 95: 3555 46. Toy PH, Newcomb M, Coon MJ, Vaz ADN (1998) J Am Chem Soc 120: 9718 47. Siegbahn PEM, Crabtree RC (1999) J Am Chem Soc 121: 117 48. Linde C, Akermark B, Norrby PO, Svensson M (1999) J Am Chem Soc 121: 5083 49. Sisemore MF, Selke M, Burstyn JN, Valentine JS (1997) Inorg Chem 36: 979 50. Selke M, Valentine JS (1998) J Am Chem Soc 120: 2652 51. Wertz DL, Sisemore MF, Selke M, Driscoll J, Valentine JS (1998) J Am Chem Soc 120: 5331 52. Ozette K, Leduc P, Palacio M, Bartoli JF, Barkigia KM, Fajer J, Battioni P, Mansuy D (1997) J Am Chem Soc 119: 6442 53. Gisdakis P, Antonczak S, KoÈstlmeier S, Herrmann WA, RoÈsch N (1998) Angew Chem Int Ed 37: 2211 54. Chen K, Que L (1999) Angew Chem Int Ed 38: 2227 55. Groves JT, Haushalter RC, Nakamura M, Nemo TE, Evans BJ (1981) J Am Chem Soc 103: 2884

34

B. Meunier á J. Bernadou

56. Meunier B, Robert A, Pratviel G, Bernadou J (2000) In: Kadish K, Smith K, Guilard R (eds) The porphyrin handbook. Academic Press, San Diego, vol 4, chap 31, pp 119±187 57. Mandon D, Weiss R, Jayaraj K, Gold A, Terner J, Bill E, Trautwein AX (1992) Inorg Chem 31: 4404 58. Gadjhede M, Schuller DJ, Henriksen A, Smith AT, Poulos TL (1997) Nature Structural Biol 4: 1032 59. Urano Y, Higuchi T, Hirobe M, Nagano T (1997) J Am Chem Soc 119: 12,008 60. Robert A, Loock B, Momenteau M, Meunier B (1991) Inorg Chem 30: 706 61. Collins TJ, Gordon-Wylie SW (1989) J Am Chem Soc 111: 4511 62. Collins TJ (1994) Acc Chem Res 27: 279 63. Bottomley LA, Neely FL (1997) Inorg Chem 37: 5435 64. Du Bois J, Tomooka CS, Hason J, Carreira EM (1997) Acc Chem Res 30: 364 65. Bortolini O, Meunier B (1983) JCS Chem Commun 1364 66. Bortolini O, Ricci M, Meunier B, Friant P, Ascone I, Goulon J (1986) Nouv J Chim 10: 39 67. Ayougou K, Bill E, Charnock JM, Garner CD, Mandon D, Trautwein AX, Weiss R, Winkler H (1995) Angew Chem Int Ed Engl 34: 343 68. Groves JT, Lee J, Marla SS (1997) J Am Chem Soc 119: 6269 69. Jin N, Groves JT (1999) J Am Chem Soc 121: 2923 70. English AM, Tsaprailis G (1995) Adv Inorg Chem 43: 79 71. Meunier B (2000) In: Meunier B (ed) Biomimetic oxidations catalyzed by transition metal complexes. Imperial College Press, London (in press) 72. Dawson JH (1988) Science 240: 433 73. Penner-Hahn JE, Eble KS, McMurry TJ, Renner M, Balch AL, Groves JT, Dawson JH, Hodgson KO (1986) J Am Chem Soc 108: 7819 74. Chang CS, Yamazaki I, Sinclair R, Khalid S, Powers L (1993) Biochemistry 32: 923 75. Paeng KJ, Kincaid JR (1988) J Am Chem Soc 110: 7913 76. Chuang WJ, Van Wart HE (1992) J Biol Chem 267: 13,293 77. Ozaki SI, Inada Y, Watanabe Y (1998) 120: 8020 78. Nakajima R, Yamazaki I (1987) J Biol Chem 262: 2576 79. Ator MA, David SK, Ortiz de Montellano PR (1987) J Biol Chem 262: 14,954 80. Chang YT, Veitch NC, Leow GH (1998) J Am Chem Soc 120: 5168 81. Gajhede M, Schuller DJ, Henriksen A, Smith AT, Poulos TL (1997) Nat Struct Biol 4: 1032 82. Klibanov AM, Berman Z, Alberti BN (1981) J Am Chem Soc 103: 6263 83. Ortiz de Montellano PR, Choe YS, DePillis G, Catalano CE (1987) J Biol Chem 262: 11,641 84. Baciocchi E, Lanzalunga O, Malandrucco S (1996) J Am Chem Soc 118: 8973 85. Grif®n BW (1991) In: Everse J, Everse KE, Grisham MB (eds) Peroxidases in chemistry and biology, vol II. CRC Press, Boca Raton, pp 85±137 86. Blanke SR, Hager LP (1988) J Biol Chem 263: 18,739 87. Sundaramoorthy M, Terner J, Poulos TL (1995) Structure 3: 1367 88. Sundaramoorthy M, Terner J, Poulos TL (1998) Chem & Biol 5: 461 89. Egawa T, Miki H, Ogura T, Makino R, Ishimura Y, Kitagawa T (1992) FEBS Let 305: 206 90. Wagenknecht HA, Woggon WD (1997) Chem & Biol 4: 367 91. Geigert J, Lee TD, Dalietos DJ, Hirano DS, Neidleman SL (1986) Biochem Biophys Res Commun 136: 778 92. Debrunner PG, Dexter AF, Schulz CE, Xia YM, Hager LP (1996) Proc Natl Acad Sci USA 93: 12,791 93. Allain EJ, Hager LP, Deng L, Jacobsen EN (1993) 115: 4415 94. Dexter AF, Lakner FJ, Campbell RA, Hager LP (1995) 117: 6412 95. Colonna S, Gaggero N, Manfredi A, Casella L, Gullotti M, Carrea G, Pasta P (1990) Biochemistry 29: 10,465 96. Colonna S, Gaggero N (1994) Phosph Sulf Silicon 95±96: 103 97. Kobayashi S, Nakano M, Kimura T, Schaap AP (1987) Biochemistry 26: 5019

Active Iron-Oxo and Iron-Peroxo Species in Cytochromes P450 and Peroxidases

35

98. Van Handel MJH, Primus JL, Teunis C, Boersma MG, Osman AM, Veeger C, Rietjens IMCM (1998) Inorg Chem Acta 275/276: 98 99. Meunier B, Guilmet E, De Carvalho ME, Poilblanc R (1985) J Am Chem Soc 106: 6668 100. Robert A, Meunier B (1988) New J Chem 12: 885 101. Nam W, Valentine JS (1993) J Am Chem Soc 115: 1772 102. The denomination ``oxo-hydroxo tautomerism'' which is reminiscent of the keto-enol tautomerism appears to be more precise than ``redox tautomerism'' that we initially proposed in [103] 103. Bernadou J, Fabiano AS, Robert A, Meunier B (1994) J Am Chem Soc 116: 937 104. Ingold KU, MacFaul PA (2000) In: Meunier B (ed) Biomimetic oxidations catalyzed by transition metals. Imperial College Press, London, chap 2 (in press) 105. Hashimoto S, Tatsuno Y, Kitagawa T (1986) Proc Natl Acad Sci USA 83: 2417 106. Makino R, Uno T, Nishimura Y, Iizuka T, Tsuboi M, Ishimura Y (1986) J Biol Chem 261: 8376 107. Hashimoto S, Tatsuno Y, Kitagawa T (1987) J Am Chem Soc 109: 8096 108. Groves JT, Stern MK (1987) J Am Chem Soc 109: 3812 109. Groves JT, Stern MK (1988) J Am Chem Soc 110: 8628 110. Balahura RJ, Sorokin A, Bernadou J, Meunier B (1997) Inorg Chem 36: 3488 111. Lee KA, Nam W (1997) J Am Chem Soc 119: 1916 112. Pitie M, Bernadou J, Meunier B (1995) J Am Chem Soc 117: 2935 113. Song R, Sorokin A, Bernadou J, Meunier B (1997) J Org Chem 62: 673 114. Yang SJ, Nam W (1998) Inorg Chem 37: 606 115. Prince S, KoÈrber F, Cooke PR, Lindsay Smith JR, Mazid MA (1993) Acta Cryst C49: 1158 116. Jeon S, Bruice TC (1992) Inorg Chem 31: 4843 117. Nam W, Goh YM, Lee YJ, Lim MH, Kim C (1999) Inorg Chem 38: 3238 118. Wietzerbin K, Muller JG, Jameton RA, Pratviel G, Bernadou J, Meunier B, Burrows CJ (1999) Inorg Chem 38: 4123 119. Sorokin A, Meunier B (1998) Eur J Inorg Chem 1269 120. Balch AL, Chan YW, Cheng RJ, La Mar GN, Latos-Grazynski L, Renner MW (1984) J Am Chem Soc 106: 7779 121. Groves JT, Gross Z, Stern MK (1994) Inorg Chem 33: 5065 122. Kaaret TW, Zhang GH, Bruice TC (1991) J Am Chem Soc 113: 4652 123. Rachlewicz K, Latos-Grazynski L (1996) Inorg Chem 35: 1136

Nucleophilicity of Iron-Peroxo Porphyrin Complexes Diana L. Wertz1, Joan Selverstone Valentine2 Department of Chemistry and Biochemistry, University of California Los Angeles, Los Angeles, CA 90095-1589 USA 1 E-mail: [email protected] 2 E-mail: [email protected]

For the past 20 years, cytochrome P450 researchers have sought to identify and to characterize the reactive intermediates in reactions of these enzymes. This review focuses on one of those postulated intermediates, the ferric heme peroxo complex, [(porphyrin)Fe(III)(O22 )]), a species which has been postulated to be formed transiently in the P450 catalytic cycle. Ferric peroxo porphyrin complexes, inorganic complexes that model the peroxo species, have been synthesized and their chemical reactivities characterized for comparison with the enzymes. Such studies have identi®ed certain peroxo porphyrins as remarkably strong nucleophiles capable of oxidizing a variety of electron-poor molecules. While the ferric heme peroxo intermediate, in the majority of P450 enzymes, rapidly converts to an oxoferryl species, some enzymes, e.g., aromatase, lanosterol 14a-demethylase, progesterone 17a-hydroxylase/17,20-lyase, and NO synthase, appear to use this intermediate as the active oxidant. Additionally, studies of ferric peroxo porphyrin complexes have increased our understanding of the nature of the P450 catalytic cycle and of the mechanisms of generation of other reactive intermediates used in P450 enzymes. Keywords: Iron-Peroxo, Superoxide, Cytochrome P450, Metal Peroxo, Oxoferryl

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

38

1.1

The P450 Catalytic Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

2

Ferric Peroxo Intermediate in P450 Systems . . . . . . . . . . . . . .

41

2.1

Ferric Peroxo Species as an Active Oxidant in Biological Systems . . . . . . . . . . . . . . . . . . . . . P450 Aromatase . . . . . . . . . . . . . . . . . . . . . . . . Cytochrome P450 2B4 and Cytochrome P450 2E1 Lanosterol P450-14a-demethylase . . . . . . . . . . . . Progesterone 17a-hydroxylase/17,20-Lyase . . . . . NO Synthase . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . .

41 42 44 45 46 47

3

Inorganic Model Complexes of the Ferric Peroxo Intermediate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

47

3.1 3.2

Synthesis of Ferric Peroxo Porphyrin Complexes . . . . . . . . . . . Characterization of Ferric Peroxo Porphyrin Complexes . . . . . .

48 49

4

Reactivity of Ferric Peroxo Porphyrin Complexes . . . . . . . . . .

49

2.1.1 2.1.2 2.1.3 2.1.4 2.1.5

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

38 4.1 4.2

D.L. Wertz á J.S. Valentine

Demonstration of the Nucleophilic Nature of Ferric Peroxo Porphyrin Complexes . . . . . . . . . . . . . . . . . . . . . . . . . . Epoxidation of Electron-Poor Enones by Ferric Peroxo Porphyrin Complexes . . . . . . . . . . . . . . . . . . . . . . . . . .

50 51

5

Nucleophilic Reactivity of the Ferric Peroxo Porphyrin Species in Biological Systems . . . . . . . . . . . . . . . . . . . . . . . . .

5.1

Ferric Peroxo Porphyrin Model Complexes as Mimics of the Third Catalytic Step of Aromatase . . . . . . . . . . . . . . . . . Ferric Peroxo Porphyrin Complexes as Mimics of Lanosterol-14a-demethylase, Progesterone 17a-hydroxylase/17,20-Lyase, and NO Synthase . . . . . . . . . . . . .

54

Failure to Convert the Synthetic Ferric Peroxo Porphyrin Complex to an Electrophilic High Valent Oxoferryl Species in a Fashion Analogous to Cytochrome P450 . . . . . . . .

55

7

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

56

8

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

58

5.2

6

53 53

List of Abbreviations CH3CN DMSO [Fe(F20TPP)O2]) [Fe(OEP)O2]) [Fe(PPIXDME)O2]) [Fe(TMP)O2]) [Fe(TPP)O2]) K222 KO2 [Mn(TPP)O2]) NOS P2) P ) PPh3 THF

Acetonitrile Dimethyl sulfoxide Peroxoferritetra(penta¯uorophenyl)porphyrin Peroxoferrioctaethylporphyrin Peroxoferriprotoporphyrin IX dimethyl ester Peroxoferritetramesitylporphyrin Peroxoferritetraphenylporphyrin Kryptate 222 Potassium superoxide Manganic peroxo tetraphenylporphyrin Nitric oxide synthase Porphyrin ligand Porphyrin cation radical formed from oxidation of the porphyrin ligand Triphenylphosphine Tetrahydrofuran

1 Introduction The cytochrome P450 enzymes comprise a large family of heme-containing enzymes with a cysteine in the active site that acts as an axial ligand to the

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

39

heme iron. To date, over 470 individual cytochrome P450 sequences have been reported [1]. Cytochrome P450 enzymes are monooxygenases, i.e., they catalyze the incorporation of one oxygen atom of dioxygen into substrate while the other oxygen atom is reduced by two electrons (usually from NADPH) to give water (Eq. 1) [2]: P450

Substrate ‡ O2 ‡ 2e ‡ 2H‡ ! Substrate…O† ‡ H2 O

…1†

There is great diversity in reactivity among the different P450 enzymes with over 40 different known types of transformations [3, 4]. The most often encountered processes are hydroxylation, epoxidation, heteratom dealkylation, heteratom oxidation, and alcohol oxidation. This diversity makes the P450 enzyme family one of the body's most important resources for the biotransformation of naturally occurring biomolecules and for the oxidative metabolism of xenobiotics. In fact, Coon et al. recently predicted that each mammalian species will be found to have up to one hundred P450 isoforms, a thousand or more inducers of P450, and a million potential substrates [3]. The ability of one enzyme family to participate in such varied transformations can be understood by examining the P450 catalytic cycle, the process by which the active oxidants are formed. As we will see, slight changes in the enzyme active site may control the formation of one active oxidant vs another and thus provide some measure of ¯exibility to the P450 family [5]. 1.1 The P450 Catalytic Cycle

The catalytic cycle of cytochrome P450 enzymes begins with the binding of substrate to the six-coordinate low-spin ferric resting form of the enzyme (Fig. 1) [2, 5]. Binding of substrate displaces the water ligand and induces a change in spin state from the low-spin resting form to the high-spin ferric form, a form in which reduction is more thermodynamically favorable. This high-spin ferric species is then reduced to the ferrous state by an external reductant, most often a closely associated P450 reductase. Binding of dioxygen to the ferrous porphyrin species generates a low-spin ferrous-dioxygen complex, [(P2))Fe(II)(O2)], or a ferric superoxide complex, [(P2))Fe(III)(O2 )]. It is unclear at this time which of the two resonance structures contributes most to the resonance hybrid of this intermediate, and it may well be that the charge distribution of this intermediate will vary depending on the speci®c active site of each individual cytochrome P450 enzyme. The catalytic cycle continues as the ferrous-dioxygen complex accepts another electron to form a ferric peroxo species, [(P2))Fe(III)(O22 )]), an intermediate which is presumably usually transient but which may, in a subset of the enzymes, serve as a nucleophilic oxidant in direct reactions with the enzyme-bound substrate (see below). The nucleophilic reactivity of the ferric peroxo species might be rationalized by considering it is analogous to

40

D.L. Wertz á J.S. Valentine

Fig. 1. A simpli®ed version of the P450 catalytic cycle. The timing of proton donation events

and axial ligand coordination probably control which species acts as the active oxidant in a particular enzymatic system

HOO) where the g1 coordination to the ferric ion is analogous to the HAO bond of deprotonated peroxide. In most cases, however, the ferric peroxo complex is assumed to be rapidly protonated to give the ferric hydroperoxo intermediate, [(P2))Fe(III)(HOO))]. Again this intermediate is usually considered to be transient in nature, although it has also at times been proposed to react directly with substrate [6, 7]. The reactivity of this species, [(P2))Fe(III) (HOO))], would be expected to be analogous to the reactivity of H2O2, with the ferric ion having been substituted for a proton; thus, the ferric hydroperoxide species would be expected to be an electrophilic oxidant rather than a nucleophilic oxidant. The delivery of a second proton to the ferric hydroperoxo species results in heterolytic oxygen-oxygen bond cleavage to give an oxoferryl porphyrin cation radical species, [(P ))Fe(IV)(O2))]+. The exact mechanism for the delivery of the two protons to the ferric peroxo intermediate is currently unknown [5]. An alternative to the sequential proton delivery pathway presented above is a concerted pathway in which the protons

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

41

are delivered contemporaneously to give the oxoferryl porphyrin cation radical species directly. A great deal of effort has been directed towards understanding the oxoferryl porphyrin cation radical intermediate because most of the transformations of the P450 enzyme family are accomplished through the reaction of substrate with this species. Many excellent reviews that discuss oxoferryl species reactivity and model complexes are available to the reader [2, 8]. The early intermediates, such as the ferric peroxo, have received less attention due in part to the general perception that these intermediates are not the primary active oxidants used in P450 catalysis. Recently this view has changed as several enzymes now appear to use early intermediates to achieve oxidative chemistry [2, 6, 7].

2 Ferric Peroxo Intermediate in P450 Systems This review will discuss efforts that have been directed towards the study of enzymes using the ferric peroxo species as the active oxidant. We will then describe attempts that have been made to model this active site species and discuss what is known about the reactivity and properties of iron peroxo porphyrin model complexes. These biological and model studies have furthered our understanding of the modes of reactivity available to the ferric peroxo species and have provided us with a framework from which the P450 catalytic cycle and P450 enzymes in general can be better understood. 2.1 Ferric Peroxo Species as an Active Oxidant in Biological Systems

While in most P450 enzymes the peroxo intermediate is rapidly converted to the oxoferryl species upon proton donation (Fig. 2), a subset of P450 enzymes have active sites in which proton donation to the peroxo species may be inhibited. In these enzymes it then becomes possible for the peroxo intermediate, with a now extended lifetime, to act as an active oxidant when

Fig. 2. The active site arrangement of a P450 enzyme where oxoferryl formation would be

favored

42

D.L. Wertz á J.S. Valentine

given the appropriate substrate. For example, one mechanism that may effectively inhibit proton donation is the disruption of hydrogen bonding between the active site water and hydroxyl group and the oxygen atoms of the peroxo by the proper placement of a substrate molecule (Fig. 3) [9]. Evidence for the ferric peroxo species as an active oxidant has been obtained through the study of wild type P450 enzymes as well as of P450 mutants in which the appropriate proton-donating residue is mutated. Several enzymes that appear to use ferric peroxo species to accomplish direct oxidations are now known; experiments with these enzymes that supported the feasibility of the ferric peroxo species as the reactive oxidizing intermediate are described below. 2.1.1 P450 Aromatase Cytochrome P450 aromatase is perhaps the most thoroughly studied of the enzymes that appear to use the iron peroxo species as an active intermediate. Interest in aromatase stems from the uniqueness of the transformation that it catalyzes, i.e., the conversion of androgens to estrogens [10]. Additionally, aromatase is studied for the biological signi®cance that it holds for humans. Studies indicate that aromatase levels are higher in breast cancer tissue vs normal breast tissue [3]. The production of estrogen may also be important for the development and maintenance of memories [11]. Aromatase catalyzes the conversion of human androgens to estrogens in three steps as shown in Fig. 4 [10]. Each transformation requires one equivalent each of NADPH and dioxygen. The ®rst step involves the hydroxylation of the C19 methyl group to give the C19 alcohol. The second step involves a second hydroxylation of the C19 carbon to give the C19 diol, which presumably loses water rapidly to become the C19 aldehyde. Both the ®rst and second steps are thought to be typical P450 type hydroxylations mediated by the oxoferryl porphyrin cation radical species. The unusual third step, however, requires the cleavage of the C19 aldehyde to give an aromatized A ring [12]. The active oxidant that performs this remarkable transformation

Fig. 3. The active site arrangement of a P450 enzyme where the ferric peroxo intermediate

would be the favored active oxidant

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

43

Fig. 4. The conversion of androgens to estrogens by P450 aromatase

has been the subject of debate among P450 researchers for some time [6, 12±15]. Proposed active oxidants for the third step have included the oxoferryl species, the ferric hydroperoxo species, and the ferric peroxo species. However, much evidence (see below) has accumulated that points to the ferric peroxo porphyrin species as the most likely candidate for the active oxidant of aromatase's third catalytic step. The ®rst researcher to propose the ferric peroxo species as the active oxidant was Akhtar et al. in 1981 [10, 12]. Through elegant isotopic labeling studies, Akhtar et al. established that one atom of the dioxygen taken up by the enzyme for the third step is incorporated into the formic acid product [12]. In the same study, the hydrogen of formic acid was shown to be the original hydrogen of the C19 aldehyde. Additional labeling studies demonstrated that the other oxygen in formic acid originated from the dioxygen required for the ®rst hydroxylation step [16]. The availability of several P450 crystal structures, including the ®rst by Poulos of P450cam [17], allowed Graham-Lorence et al. to construct a threedimensional hypothetical model of the aromatase active site [9]. One active site residue that they identi®ed as being potentially important in determining the mechanistic pathway of the enzyme is Thr310. That threonine residue is positioned in the active site so that hydrogen bonding to the oxygen of the C19 aldehyde is possible. This hydrogen bonding increases the electrophilicity of the aldehyde. The fact that Thr310 may function to activate the substrate suggests that proton donation to the ferric peroxo species may be prohibited. Graham-Lorence et al. suggest that this active site setup may serve to increase the nucleophilicity of the ferric peroxo species and prevent formation of the ferric hydroperoxo intermediate and its subsequent degradation to the oxoferryl species. One potential mechanism for the third step of aromatase was proposed by Graham-Lorence et al. and is shown in Fig. 5 [9].

44

D.L. Wertz á J.S. Valentine

Fig. 5. One potential mechanism of the third step in the conversion of testosterone to

estradiol by P450 aromatase

Akhtar, Groves and others propose that the ferric peroxo species attacks the C19 aldehyde to give an iron peroxohemiacetal intermediate [6, 8, 18]. The peroxohemiacetal may then decompose in a concerted or stepwise fashion [19, 20]. Although the degradation pathway of the acetal intermediate is not known, Townsley has established that the 1b (b = above) proton is selectively removed and that enolization of the a,b-unsaturated ketone moiety is necessary in order to produce a fully aromatized steroid A ring [21]. Substrate analogs made and studied by Oh and Robinson suggest that enolization prior to deformylation is favorable [18]. Additionally, the active site modeling studies of Graham-Lorence et al. indicate that Asp309 is in close enough proximity to the C2 protons to act as a base capable of generating the enolized A ring [9]. 2.1.2 Cytochrome P450 2B4 and Cytochrome P450 2E1 Convincing evidence for the existence of an iron peroxo intermediate as an active oxidant in biological systems has arisen from mutagenesis studies of several microsomal P450 enzymes [3]. Cytochrome P450 2B4 and cytochrome P450 2E1 are members of the P450 family that are able to hydroxylate substrates as well as deformylate aldehydes [22±26]. Studies of P450 2B4 and P450 2E1 mutants have provided evidence for the use of the peroxo and hydroperoxo intermediates if protons in the active site are not available to promote OAO bond cleavage [7]. In the enzyme P450 2B4, an active site threonine residue is believed to act as a proton donor. Mutation of that residue to alanine (T302A) resulted in greatly decreased hydroxylation of substrates and a ®ve- to tenfold increase in deformylation of cyclohexanecarboxaldehyde through what is believed to be the peroxo intermediate [25]. Vaz et al. had shown in 1994 that microsomal P450 2B4 is able to aromatize an isotopically labeled truncated version of the

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

45

aromatase natural substrate with selective removal of the 1b hydrogen, as is seen in P450 aromatase [22]. Therefore, it appears that P450 2B4 has an ability to act in a manner similar to aromatase. Removal of the Thr302 active site residue presumably shifts the active oxidant to predominately the ferric peroxo species and, therefore, deformylation is greatly enhanced. Remarkably, the T302A mutation of P450 2E1 resulted in increased epoxidative reactivity towards styrene and decreased allylic hydroxylation activity towards cyclohexene and butene [7]. In cytochrome P450 2E1, trends in the ratios of the rates of epoxidation and hydroxylation support the conclusion that the active oxidant is the ferric hydroperoxo-iron species. Therefore, the removal of the Thr302 amino acid active site proton may shift the protonation equilibrium towards the ferric peroxo intermediate in the case of P450 2B4 and towards a ferric hydroperoxide intermediate in the case of P450 2E1. The studies discussed above illustrate the idea that an early iron peroxo species will have a suf®cient lifetime to act as an active oxidant if proton delivery to the oxygens of the peroxo is inhibited [7, 27±29]. It is important to note that many other researchers have studied mutated P450 enzymes in order to understand how active site proton donating ability affects reactivity; however, we have only presented studies which have direct relevance to the mechanism of aromatase [28±30]. Additionally, Sligar et al. have recently reported the spectroscopic characterization of a reduced ferrous dioxygen species in P450 cam [31]. Density functional calculations by Waskell et al. corroborate the assignment of the observed species as the peroxo intermediate [32]. 2.1.3 Lanosterol P450-14a-demethylase The enzyme lanosterol-14a-demethylase performs a three step transformation of lanosterol that is reminiscent of the transformation performed by aromatase (Fig. 6) [4, 33]. Consequently, it is tempting to assume the mechanisms are similar [6]. Like aromatase, the ®rst two steps in the oxidation of lanosterol are typical P450-type hydroxylations. It is at this point, however, that the two

Fig. 6. The deformylation of lanosterol by lanosterol 14a-demethylase

46

D.L. Wertz á J.S. Valentine

enzymes diverge mechanistically. Instead of direct attack of the ferric peroxo complex onto the aldehyde and direct decomposition as is seen in aromatase, lanosterol-14a-demethylase probably forms an ester intermediate as shown in Fig. 7 [34]. There is much support for the proposal that the third step involves a ferric peroxo mediated Baeyer-Villiger type reaction. The ester shown in Fig. 7 has been isolated and has been shown to proceed to the ole®n product in the presence of lanosterol 14a-demethylase [35]. 2.1.4 Progesterone 17a-Hydroxylase/17,20-Lyase The enzyme progesterone 17a-hydroxylase/17,20-lyase catalyzes the cleavage of the C17 side chain of progesterone (or pregnenolone) to form androstenedione (or dehydroepiandrostene) [33]. This unique enzyme is able to follow a hydroxylase/lyase pathway (as shown by the top pathway in Fig. 8) or an aldehyde cleavage pathway (as shown at the bottom of Fig. 8). Recent studies by Akhtar et al. [36] and by Swinney and Mak [37] used isotopic labeling to establish that the reaction does not proceed via an oxoferryl species as the

Fig. 7. The biological Baeyer-Villiger type reaction proposed as a part of the mechanism for

the deformylation of the D ring of lanosterol by lanosterol 17a-demethylase

Fig. 8. The two potential oxidative pathways catalyzed by progesterone 17a-hydroxylase/

17,20-lyase

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

47

active oxidant. Most evidence supports a ferric peroxo porphyrin species being the active oxidant in both pathways. In the hydroxylation/lyase sequence, the substrate is hydroxylated alpha to the carbonyl of the ketone by an oxoferryl species, as is typical of hydroxylation reactions. In the lyase step, however, the ferric peroxo porphyrin is proposed to attack the ketone moiety to form a peroxyhemiketal type intermediate [38]. The mode of decomposition of this intermediate is currently unknown, but it is most likely radical in nature (although a Baeyer-Villiger route is also possible) [39]. A minor pathway as shown on the bottom of Fig. 8 is also possible if the peroxyhemiketal intermediate forms before hydroxylation [4]. 2.1.5 NO Synthase NO synthase (NOS) catalyzes the ®ve-electron oxidation of L -arginine to and nitric oxide, as shown in Fig. 9 [8]. Spectroscopic evidence presented by Degray et al. [40] and Harris et al. [41] suggested that NOS can be classi®ed as a P-450 enzyme. The transformation of L -arginine to L -citrulline is divided into two separate steps. Crane proposes that in the ®rst oxidative step a ferric peroxo porphyrin intermediate removes a proton from a guanidinium moiety of L -arginine [45]. Upon protonation of the ferric peroxo porphyrin complex, the intermediate then decomposes to an oxoferryl cation radical porphyrin species which is then capable of hydroxylating the nitrogen of the guanidinium. Before the next step, an electron transfer must occur, presumably from the substrate to the enzyme [8]. Then, in the second step, it is proposed that a ferric peroxo porphyrin species attacks the carbon of the urea oxime moiety as shown in Fig. 10 [43]. At this time, the mechanism and active oxidant identity during each transformation are subjects of intense debate, but the recent crystal structures of NOS coupled with new research ®ndings may help resolve these issues [42, 44]. L -citrulline

3 Inorganic Model Complexes of the Ferric Peroxo Intermediate Researchers have been actively studying ferric peroxo porphyrin model complexes for over 20 years because of their relevance to the above enzymatic

Fig. 9. The conversion of L -arginine to L -citrulline which is catalyzed by NO synthase

48

D.L. Wertz á J.S. Valentine

Fig. 10. One mechanism that has been proposed for the carbon-nitrogen bond cleavage step

catalyzed by NO synthase

transformations. This section covers the synthesis and characterization of various ferric peroxo porphyrin complexes. In later sections, we review characterization of the reactivity of ferric peroxo porphyrin complexes as well as consider studies which attempt to model the basic transformations that occur in the above-mentioned enzymes. 3.1 Synthesis of Ferric Peroxo Porphyrin Complexes

The synthesis of a ferric peroxo porphyrin complex, [Fe(TPP)O2]), was ®rst reported in 1978 [45, 46]. Valentine et al. reported a preparation that took advantage of the solubility of potassium superoxide in the presence of crown ethers [47]. The synthesis, as shown in Fig. 11, is similar to the P450 catalytic cycle in that it provides two electrons and a dioxygen molecule to give an iron(III) peroxo porphyrin. The synthesis is possible only in aprotic solvents such as acetonitrile and toluene due to the disproportionation of superoxide in protic solvents. Khenkin et al. also reported the synthesis of a compound with the spectroscopic properties of the ferric peroxo porphyrin complex, but it was assumed to be a ferrous superoxo porphyrin complex at the time [46]. Another synthetic route to ferric peroxo porphyrin complexes was reported in 1981 when Welborn et al. [48] reported the one-electron reduction of a porphyrin ferrous dioxygen complex to give a ferric peroxo porphyrin complex. A third synthetic route reported by Reed in 1982 illustrated that

Fig. 11. The synthesis of ferric porphyrin peroxo complexes is possible by the addition of

two equivalents of superoxide to a ferric porphyrin complex. The ®rst equivalent of superoxide reduces the ferric porphyrin to the ferrous porphyrin. Oxidative addition of a superoxide molecule to the ferrous porphyrin produces the ferric g2-peroxo porphyrin complex

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

49

dioxygen addition to an iron(I) porphyrin compound also yielded the ferric peroxo porphyrin complex [49]. 3.2 Characterization of Ferric Peroxo Porphyrin Complexes

The [Fe(OEP)O2]) complex was reported in 1980, along with spectroscopic characterization of both [Fe(OEP)O2]) and [Fe(TPP)O2]) [47]. In order to establish whether the peroxide ligand was bound in an g1 or an g2 fashion, the frequency of the OAO bond in [Fe(OEP)O2]) was measured by IR spectroscopy. g1-Superoxo type metal complexes typically exhibit oxygenoxygen bond frequencies in the 1030±1180 cm)1 region while those of g2-peroxo type metal complexes are in the 750±950 cm)1 region [50]. The 806 cm)1 vibration observed for [Fe(OEP)O2]), along with the observed shift in frequency to 759 cm)1 upon substitution with 18O2, suggest that the actual structure is a ferric g2-peroxo complex [47]. Further evidence for a g2peroxo type metal complex is provided by the EPR spectrum of [Fe(OEP)O2]). Signals at g = 4, g = 8 (weak), and g = 2 (weak) indicate the presence of a high spin rhombic ferric complex. The observed rhombicity supports the hypothesis that the peroxo ligand in [Fe(OEP)O2]) is bound to iron in an g2 fashion. The deuterium NMR characterization of the [Fe(TPP)O2]) complex by Shirazi and Goff is also consistent with a high spin ferric species [51]. Although a crystal structure of a ferric peroxo porphyrin is still not available, EXAFS studies on [Fe(TPP)O2]) are consistent with a peroxide bound in an g2 fashion [52]. Additionally, the [Mn(TPP)O2]) complex has been crystallized and is known to contain a peroxo ligand bound to manganese in an g2 fashion, as shown in Fig. 12 [53]. Mossbauer, EPR, and magnetic susceptibility data for [Fe(TPP)O2]) are all consistent with a geometry for the ferric peroxo complex that is similar to that observed for [Mn(TPP)O2]) [54]. Therefore, the crystal structure of the manganese peroxo porphyrin probably re¯ects the nature of the ferric peroxo porphyrin as well.

4 Reactivity of Ferric Peroxo Porphyrin Complexes After the initial synthesis of [Fe(TPP)O2]), many additional ferric peroxo porphyrin complexes were synthesized and studied in order to determine whether they exhibited nucleophilic or electrophilic properties. Attempts to react ferric peroxo porphyrins with electron-rich substrates such as styrene and triphenylphosphine have resulted in little or no reactivity. For example, Welborn et al. [48] observed no epoxidation of styrene while a recent study showed virtually no oxidation of triphenylphosphine in THF [55, 56]. These studies established that ferric peroxo porphyrin complexes are not, by nature, electrophilic. The attention of researchers then shifted in order to explore the potential for nucleophilic reactivity.

50

D.L. Wertz á J.S. Valentine

Fig. 12. The crystal structure of [Mn(TPP)O2]) complex as determined by Van Atta et al.

[53] The associated cation, potassium K222, has been removed for this ®gure

4.1 Demonstration of the Nucleophilic Nature of Ferric Peroxo Porphyrin Complexes

The ®rst demonstration of the nucleophilic nature of a ferric peroxo complex was in 1982 when Khenkin and Shteinmann reacted [Fe(TPP)O2]) with acetic anhydride or acetyl chloride and observed the oxidation of cyclohexane to cyclohexanol [57]. They postulated that the ferric peroxo porphyrin complex reacts with the acyl anhydride at )70 °C to form, presumably, a peracetic acid iron(III) porphyrin complex. This species then degrades above )70 °C to give the high valent iron oxo species which is capable of alkane hydroxylation. Khenkin and Shteinmann later published a full report on the subject in 1984 and provided further evidence to support the formation of the peracetic acid iron(III) porphyrin complex [58]. Shappacher and Weiss [59] and Groves and Watanabe [60] observed additional examples of ferric peroxo porphyrins reacting with electrophiles to produce iron(IV) oxo porphyrin species. In 1985 Schappacher and Weiss

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

51

demonstrated that ferric peroxo picket fence porphyrins react with carbon dioxide to give an iron(IV) oxo porphyrin adduct [59]. In this case a monoperoxy carbonate iron(III) porphyrin intermediate was not detected but presumably formed. Groves and Watanabe also reported the reaction of ferric peroxo complexes with acid anhydrides and acyl halides to produce the iron(IV) oxo porphyrin adduct [60]. Therefore, by 1986 it was established that ferric peroxo complexes could be activated by addition of an acyl group, and presumably a proton, to give high valent iron oxo species. This type of reaction is similar to protonation of the ferric peroxo porphyrin and subsequent degradation of the hydroperoxo species as seen in Fig. 1. It is also possible to generate lower valent porphyrin products by reacting the ferric peroxo complex with electrophiles. For example, Mitzkal et al. determined that [Fe(TPP)O2]) reacted with SO2 to form the [Fe(II)(TPP)] (DMSO)2 complex as the porphyrin product [61]. SO24 was presumed to be the primary product of SO2 oxidation, although, the authors actually observed S2O27 , a product of sulfate condensation. These early studies of ferric peroxo porphyrin reactivity were important for several reasons. First, the ferric peroxo porphyrin could now be classi®ed as a nucleophilic oxidant. Second, the peroxyacetate ferric porphyrin intermediate generated in the reaction of the ferric peroxo porphyrin with acyl anhydrides and acetic anhydride was also of interest to researchers because oxygenoxygen bond cleavage was shown to occur to give higher valent iron oxo porphyrin compounds. While these studies were the ®rst models of the events that occur in P450 actives sites to generate the oxoferryl intermediate, they did not necessarily help to establish the ferric peroxo species as a viable active oxidant in biological systems. Later work would provide support for the idea that the ferric peroxo species may be an important active oxidant of electrophilic substrates in certain select P450 enzymes. Several of the studies that helped to establish the viability of the ferric peroxo porphyrin species as a primary oxidant are next considered. 4.2 Epoxidation of Electron-Poor Enones by Ferric Peroxo Porphyrin Complexes

An important step towards establishing the oxidizing ability of ferric peroxo porphyrin complexes came when Sisemore et al. demonstrated that the very nucleophilic [Fe(TMP)O2]) epoxidized electron-de®cient ole®ns [62]. The epoxidation, shown in Fig. 13, was the ®rst demonstration of direct oxygen insertion into an ole®nic bond by a ferric peroxo porphyrin complex. In the proposed mechanism, the authors suggest that the ferric peroxo species attacks the electron-de®cient b-carbon of the enone to yield an epoxide product. This process is similar to the epoxidation of certain alkenes by basic hydrogen peroxide. In a later study, Selke et al. [63] synthesized and characterized [Fe(PPIXDME)O2]), the ferric peroxo complex of the biological porphyrin. The [Fe(PPIXDME)O2]) was found to be as effective at oxygen transfer to

52

D.L. Wertz á J.S. Valentine

Fig. 13. The epoxidation of menadione by [Fe(TMP)O2]) in CH3CN

menadione as [Fe(TMP)O2]) (75% conversion to epoxide). This study established that the heme group used in most P450 enzymes is capable of affecting direct nucleophilic attack on substrates in the ferric peroxo form. Remarkably, in the same study triphenylphosphine was found to bind weakly to the electron-de®cient ferric peroxo complex [Fe(F20TPP)O2]). This weak binding is reminiscent of the axial ligation that occurs in P450 active sites by the active site cysteinyl thiolate. Selke and Valentine [64] later demonstrated the importance of axial ligation in activating ferric peroxo porphyrin species when they reported the ``switching-on'' of the formerly unreactive [Fe(F20TPP)O2]). In acetonitrile, [Fe(F20TPP)O2]) was unreactive towards menadione, but epoxidation occurred when the reaction was carried out in DMSO, a solvent capable of axial ligation. As shown in Fig. 14, the yield of epoxidation increased from 0% to 75%. This result suggests the possibility that the nucleophilicity of a ferric peroxo porphyrin intermediate may be enhanced in an enzyme's active site by coordination of the thiolate side chain of cysteine.

Fig. 14. Comparison of the extent of epoxidation of menadione by [Fe(F20TPP)O2]) in

acetonitrile vs DMSO. Epoxidation in DMSO may be explained by the ability of DMSO to bind as an axial ligand to the iron and increase the nucleophilic character of the peroxo oxygens

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

53

5 Nucleophilic Reactivity of the Ferric Peroxo Porphyrin Species in Biological Systems The above studies served to establish the nucleophilic nature of ferric peroxo porphyrin complexes. Since they also demonstrated that ferric peroxo porphyrins could react with biologically known functional groups, the studies established the ferric peroxo as a biologically viable P450 active oxidant. The next section considers how ferric peroxo porphyrin model complexes have been used to investigate the mechanisms of enzymes thought to react via the ferric peroxo species. 5.1 Ferric Peroxo Porphyrin Model Complexes as Mimics of the Third Catalytic Step of Aromatase

Model studies aimed at determining the mechanism of aromatase have helped researchers arrive at an accepted catalytic mechanism. In 1990, Watanabe reported that the natural substrate of aromatase reacted with an oxoferryl porphyrin to give oxidation of the aldehyde to the carboxylic acid rather than aromatization of the A ring [65]. This important result allows us to conclude that an oxoferryl species is not likely as the active oxidant of aromatase. A later study by Goto et al. [66] demonstrated that ferric porphyrin peroxo complexes, such as [Fe(TMP)O2]), were mediators of carbon-carbon bond cleavage in aliphatic aldehydes (Fig. 15). This reaction mimicked the ®rst part of the aromatization process catalyzed by aromatase and thus provided support for the concept of a ferric peroxo species as an active oxidant. Cole and Robinson obtained additional evidence for the peroxo intermediate as an active oxidant when they reported that the enolized version of the natural substrate reacted slowly with hydrogen peroxide to give the aromatized product [14, 19]. This chemical model of aromatase also supported the importance of enolization of the substrate prior to deformylation as the unenolized derivative of the natural substrate failed to react with hydrogen peroxide. In a recent model study, Wertz et al. [67] were able to show that [Fe(TMP)O2]) reacts rapidly with an enolized mimic of the A ring of androstenedione to give an aromatized product (Fig. 16A). In this reaction

Fig. 15. The carbon-carbon bond cleavage of phenylpropionaldehyde by [Fe(TMP)O2]) in

acetonitrile

54

D.L. Wertz á J.S. Valentine

Fig. 16A,B. Comparison of the reactivity of a ferric peroxo porphyrin with the natural

substrate of aromatase shown in B vs an enolized mimic of the A and B ring. The epoxidation observed in B is explained by the greater electrophilicity of the b-carbon compared to the carbonyl carbon of the aldehyde

both carbon-carbon bond cleavage and aromatization of the A ring mimic occur to yield a product analogous to what is observed in the third step of aromatase. Reaction with the non-enolized natural substrate resulted in epoxidation of the a,b-unsaturated ketone of the A ring rather than deformylation (Fig. 16B). This ®nding is consistent with Cole and Robinson's observation that only enolized mimics react with hydrogen peroxide to give deformylation [19]. Calculations by Graham-Lorence et al. also suggest that enolization is achieved in the enzymatic active site before deformylation occurs [9]. Taken together these results lend strong support for the ferric peroxo porphyrin species as the active oxidant in the third catalytic step of aromatase. 5.2 Ferric Peroxo Porphyrin Complexes as Mimics of Lanosterol-14a-demethylase, Progesterone 17a-hydroxylase/17,20-Lyase, and NO Synthase

Theoretically, it should be possible for ferric peroxo porphyrin complexes to mimic the applicable steps in lanosterol-14a-demethylase and progesterone 17a-hydroxylase/17,20-lyase. In the case of lanosterol-14a-demethylase, a ferric peroxo porphyrin complex would have to react with an aldehydecontaining molecule to give a formyl ester product, presumably via a BaeyerVilliger type mechanism. Researchers have turned to the use of ketones to explore the potential of the peroxo complex for Baeyer-Villiger chemistry because, in recent studies, aldehydes have reacted to yield deformylated or carbon-carbon bond cleaved products[66, 67]. Studying ketone reactivity has the additional bene®t of being an appropriate functional group for the study of the mechanism of progesterone 17a-hydroxylase/17,20-lyase. Direct nucleophilic attack of a ferric peroxo porphyrin complex onto the carbonyl group of a ketone and subsequent rearrangement should be within

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

55

the range of chemistry possible for the ferric peroxo. However, reactivity studies involving ketone oxidation by ferric peroxo porphyrin complexes have not resulted in Baeyer-Villiger type chemistry. Goto et al. observed carboncarbon bond cleavage, not ester formation, when 2-phenyl-butyrophenone was reacted with a ferric peroxo porphyrin complex [66]. Additionally, reactivity studies with [Fe(TMP)O2]) and various simple ketones did not result in ester formation even with strained ketones such as cyclobutanone, a ketone that would be expected to ring open rapidly to form a stable ®ve-membered ring lactone [68]. The reason for the lack of Baeyer-Villiger type chemistry remains puzzling at this time. One might postulate that the C and D ring of the steroid and active-site energetic factors may convey an unknown thermodynamic driving force that is not present in the model system. Future investigations of ferric peroxo porphyrin reactivity with substrate-like ketones and a-hydroxyketones might yield functional mimics of lanosterol-14a-demethylase and progesterone 17a-hydroxylase/17,20-lyase. The use of ferric peroxo model complexes to model the ®rst and second steps of NO synthase may yield interesting results. The carbon of the urea oxime in N-hydroxy-L -arginine should be electrophilic enough to encourage nucleophilic attack of a ferric peroxo model complex. We are unaware of any studies that use ferric peroxo porphyrin complexes to model either of the pertinent steps of NO synthase.

6 Failure to Convert the Synthetic Ferric Peroxo Porphyrin Complex to an Electrophilic High Valent Oxoferryl Species in a Fashion Analogous to Cytochrome P450 The ferric heme peroxo intermediate, [(P2))Fe(III)(O22 )]), in the cytochrome P450 mechanism is protonated twice and then undergoes heterolytic OAO bond cleavage to form the oxoferryl porphyrin cation radical species, [(P ))Fe(IV)(O2))]+. We know from the X-ray crystallography of P450 enzymes that two hydrogen bond donors, a water molecule and a hydroxyl group from an amino acid side chain, are appropriately positioned to interact with the peroxo ligand of the intermediate when it is formed and that OAO bond heterolysis occurs quickly and without the formation of observable intermediates such as the hydroperoxo species [17]. As ferric peroxo complexes seem to be able to mimic some aspects of P450 biological reactivity, one might also expect that it would be possible to observe oxygen-oxygen bond heterolysis of synthetic ferric peroxo complexes upon the addition of two protons. However, attempts to mimic this seemingly straightforward enzymatic step by protonating the synthetic ferric porphyrin peroxo complexes with a wide variety of protic and Lewis acids of varying strengths and trapping the oxoferryl intermediate with ole®ns have failed utterly [69]. Dissociation of hydrogen peroxide apparently is favored over OAO bond cleavage. In this section, we consider why we have been unable to model this critical step of the enzymatic chemistry using the synthetic porphyrins.

56

D.L. Wertz á J.S. Valentine

The synthetic ferric peroxo complexes are high spin, S = 5/2, and the peroxo ligand is thought to be coordinated in an g2 fashion. In order for these complexes to achieve OAO bond heterolysis, it is necessary for the peroxo complex to be protonated, to isomerize to a monodentate con®guration, and to undergo a spin conversion as the low spin oxoferryl porphyrin cation radical species is formed (Fig. 17, pathway A). Our failure to induce OAO bond heterolysis by protonating the synthetic ferric porphyrin peroxo complex suggests to us that this pathway is not facile and is not used by the enzyme. Deoxy ferrous heme proteins and models are high spin (S = 2), but they react with dioxygen to give low spin oxy complexes (either low spin ferrousdioxygen or low spin ferric superoxide complexes). An alternative possibility to pathway A is conversion from the low spin oxy heme intermediate, ferrousdioxygen or ferric superoxide, to the low spin oxoferryl porphyrin cation radical species via a pathway that does not require high spin intermediates. Such a pathway is feasible if the low spin oxy heme intermediate is reduced and protonated simultaneously, i.e., via proton-coupled electron transfer (pathway B). The mechanism described above in pathway B is analogous to that proposed by Babcock and coworkers for heterolytic OAO bond cleavage in cytochrome c oxidase [70]. In that case, hydrogen atom transfer from a tyrosine side chain to the oxy heme intermediate is concerted with OAO bond cleavage to produce the oxoferryl species [(P2))Fe(IV)(O2))] and the tyrosyl radical, which is analogous to the intermediate proposed for cytochrome P450, [(P ))Fe(IV)(O2))]+, except that the electron de®ciency is in the tyrosine radical rather than in the porphyrin radical. A high spin ferric peroxo complex need never be invoked.

7 Conclusions Unlike the early transition metal peroxo complexes, ferric peroxo porphyrin complexes are nucleophilic. Their reactivity with anhydrides, acyl halides, carbon dioxide, and sulfur dioxide is well documented. The electronic characteristics of the porphyrin macrocycle in¯uence the degree to which the peroxo moiety acts as a nucleophile. For example, electron-rich [Fe(TMP)O2]) epoxidizes menadione while the electron-poor peroxo complex [Fe(F20TPP)O2]) does not. Certain small molecules such as DMSO and PPh3 may serve as axial ligands to the model complexes and increase nucleophilic reactivity. This axial ligation may mimic the effect that the active site thiolate has on the nucleophilicity of the peroxo intermediate in the enzyme. c Fig. 17. Two potential pathways leading to the heterolytic oxygen-oxygen bond cleavage of a

ferric superoxide species. Pathway A is the route a synthetic ferric peroxo porphyrin complex would be expected to traverse as it is converted to the oxoferryl porphyrin cation radical intermediate. Pathway B depicts the expected spin state changes in the case of proton-coupled electron transfer

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

57

58

D.L. Wertz á J.S. Valentine

Although the ferric peroxo species was once thought to be only an intermediate necessary for the generation of the oxoferryl porphyrin cation radical species, biological and model studies have established the importance of the ferric peroxo as an active oxidant. Reactivity studies of ferric peroxo porphyrin complexes with biologically-relevant electrophiles have provided mechanistic information concerning the use of the peroxo species by P450 aromatase. The continued study of peroxo porphyrin model complexes should eventually lead to a greater understanding of the role that peroxo species play in other P450 enzymes such as lanosterol-14a-demethylase, progesterone 17ahydroxylase/17,20-lyase, and NO synthase.

8 References 1. 2. 3. 4. 5.

6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27.

Information obtained from P-450 sequence databank Sono M, Roach MP, Coulter ED, Dawson JH (1996) Chem Rev 96: 2841 Coon MJ, Vaz ADN, Bestervelt LL (1996) Faseb J 10: 429 Ortiz de Montellano PR (1995) Oxygen activation and reactivity. In: Ortiz de Montellano PR (ed) Cytochrome P-450, structure, mechanism and biochemistry, 2nd edn. Plenum Press, New York, p 245 Mueller EJ, Loida PJ, Sligar SG (1995) Twenty-®ve years of P450cam research: mechanistic insights into oxygenase catalysis. In: Ortiz de Montellano PR (ed) Cytochrome P-450, structure, mechanism and biochemistry. 2nd edn. Plenum Press, New York, p 83 Akhtar M, Lee-Robichaud P, Akhtar ME, Wright JN (1997) J Steroid Biochem Molec Biol 61: 127 Vaz ADN, McGinnity DF, Coon MJ (1998) Proc Natl Acad Sci USA 95: 3555 Groves JT, Han Y-Z (1995) Models and mechanisms of cytochrome P450 action. In: Ortiz de Montellano PR (ed) Cytochrome P-450, structure, mechanism and biochemistry, 2nd edn. Plenum Press, New York, p 3 Graham-Lorence S, Amarneh B, White RE, Peterson JA, Simpson ER (1995) Protein Science 4: 1065 Akhtar M, Calder MR, Corina DL, Wright JN (1981) J Chem Soc, Chem Comm 3: 129 Fisher CR, Graves KH, Parlow AF, Simpson ER (1998) Proc Natl Acad Sci USA 95: 6965 Akhtar M, Calder MR, Corina DL, Wright JN (1982) Biochem J 201: 569 Morand P, Williamson DG, Layne DS, Lompa-Krzymien L, Salvador J (1975) Biochemistry 14: 635 Cole PA, Robinson CH (1986) J Chem Soc, Chem Comm 1651 Korzekwa KR, Trager WF, Smith SJ, Osawa Y, Gillette JR (1991) Biochemistry 30: 6155 Stevenson DE, Wright JN, Akhtar M (1988) J Chem Soc, Perkin Trans 1: 2043 Poulos TL, Finzel BC, Howard AJ (1987) J Mol Biol 195: 687 Oh SS, Robinson CH (1993) J Steroid Biochem Mol Biol 44: 389 Cole PA, Robinson CH (1991) J Am Chem Soc 113: 8130 Watanabe Y, Ishimura Y (1989) J Am Chem Soc 111: 8047 Townsley JD, Broodie HJ (1968) Biochemistry 7: 33 Vaz ADN, Kessell KJ, Coon MJ (1994) Biochemistry 33: 13,651 Raner GM, Chiang EW, Vaz ADN, Coon MJ (1997) Biochemistry 36: 4895 Roberts ES, Vaz ADN, Coon MJ (1991) Proc Natl Acad Sci USA 88: 8963 Vaz ADN, Pernecky SJ, Raner GM, Coon MJ (1996) Proc Natl Acad Sci USA 93: 4644 Peng H-M, Raner GM, Vaz ADN, Coon MJ (1995) Arch Biochem Biophys 318: 333 Imai M, Shimada H, Watanabe Y, Matsushima-Hibiya Y, Makino R, Koga H, Horiuchi T, Ishimura Y (1989) Proc Natl Acad Sci USA 86: 7823

Nucleophilicity of Iron-Peroxo Porphyrin Complexes

28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65.

59

Raag R, Martinis SA, Sligar SG, Poulos TL (1991) Biochemistry 30: 11,420 Counts Gerber NC, Sligar SG (1994) J Biol Chem 269: 4260 Poulos TL, Raag R (1992) FASEB J 6: 674 Benson DE, Suslick KS, Sligar SG (1997) Biochemistry 36: 5104 Harris D, Loew G, Waskell L (1998) J Am Chem Soc 120: 4308 Goodman LS, Gilman A (1996) The pharmacological basis of therapeutics, 9th edn. McGraw-Hill, New York Shyadehi AZ, Lamb DC, Kelly SL, Kelly DE, Schunck W-H, Wright JN, Corina D, Akhtar M (1996) J Biol Chem 271: 12,445 Fischer RT, Trzaskos JM, Magolda RL, Lo SS, Brosz CS, Larsen B (1991) J Biol Chem 266: 6124 Akhtar M, Corina D, Miller S, Shyadehi AZ, Wright JN (1994) Biochemistry 33: 4410 Swinney DC, Mak AY (1994) Biochemistry 33: 2185 Lee-Robichaud P, Shyadehi AZ, Wright JN, Akhtar ME, Akhtar M (1995) Biochemistry 34: 14,104 Lee-Robichaud P, Akhtar ME, Akhtar M (1998) Biochemical J 330: 967 Degray JA, Lassmann G, Curtis JF, Kennedy TA, Marnett LJ, Eling TE, Mason RP (1992) J Biol Chem 267: 23,583 Harris RZ, Newmyer SL, Ortiz de Montellano PR (1993) J Biol Chem 268: 1637 Crane BR, Arvai AS, Ghosh DK, Wu C, Getzoff ED, Stuehr DJ, Tainer JA (1998) Science 279: 2121 Pufahl RA, Wishnok JS, Marletta MA (1995) Biochemistry 34: 1930 Crane BR, Arvai AS, Gachhui R, Wu C, Ghosh DK, Getzoff ED, Stuehr DJ, Tainer JA (1997) Science 278: 425 Valentine JS, McCandlish E (1978) Reactions of superoxide with metalloporphyrins. In: Dutton P, Leigh JS, Scarpa A (eds) Frontiers of biological energetics. Academic Press, New York, p 933 Kol'tover VK, Koifman OI, Khenkin AM, Shteinman AA (1982) Izvest Akad Nauk SSSR, Ser Khim 7: 1690 McCandlish E, Miksztal AR, Nappa M, Sprenger AQ, Valentine JS, Stong JD, Spiro TG (1980) J Am Chem Soc 102: 4268 Welborn CH, Dolphin D, James BR (1981) J Am Chem Soc 103: 2869 Reed CA (1982) Iron(I) and iron(IV) porphyrins. In: Kadish KM (ed) Electrochemical and spectrochemical studies of biological redox components. American Chemical Society, Washington DC, p 333 Ahmad S, McCallum JD, Shiemke AK, Appelman EH, Loehr TM, Sanders-Loehr J (1988) Inorganic Chemistry 27: 2230 Shirazi A, Goff HM (1982) J Am Chem Soc 104: 6318 Friant R, Goulon J, Fischer J, Ricard L, Schappacher M, Weiss R, Momenteau M (1985) Nouv J Chim 9: 33 Van Atta RB, Strouse CE, Hanson LK, Valentine JS (1987) J Am Chem Soc 109: 1425 Burstyn JN, Roe JA, Miksztal AR, Shaevitz BA, Lang G, Valentine JS (1988) J Am Chem Soc 110: 1382 Burstyn JN (1986) Ph.D. Thesis, University of California, Los Angeles Sisemore MF, Selke M, Burstyn JN, Valentine JS (1997) Inorganic Chemistry 36: 979 Khenkin AM, Shteinman AA (1982) Kinet Katal 23: 219 Khenkin AM, Shteinman AA (1984) J Chem Soc, Chem Comm 18: 1219 Schappacher M, Weiss R (1985) J Am Chem Soc 107: 3736 Groves JT, Watanabe Y (1986) J Am Chem Soc 108: 7834 Miksztal AR, Valentine JS (1984) Inorganic Chemistry 23: 3548 Sisemore MF, Burstyn JN, Valentine JS (1996) Angew Chem Int Ed Engl 35: 206 Selke M, Sisemore MF, Valentine JS (1996) J Am Chem Soc 118: 2008 Selke M, Valentine JS (1998) J Am Chem Soc 120: 2652 Watanabe Y, Takehira K, Shimizu M, Hayakawa T, Orita H (1990) J Chem Soc, Chem Comm 13: 927

60

D.L. Wertz á J.S. Valentine

66. Goto Y, Wada S, Morishima I, Watanabe Y (1998) J Inorg Biochem 69: 241 67. Wertz DL, Sisemore MF, Selke M, Driscoll J, Valentine JS (1998) J Am Chem Soc 120: 5331 68. Wertz DL (1999) Ph.D. Thesis, University of California, Los Angeles 69. Valentine Lab, unpublished results 70. Proshlyakov DA, Pressler MA, Babcock GT (1998) Proc Natl Acad Sci USA 95: 8020

Characterization of High-Valent Oxo-Metalloporphyrins Yoshihito Watanabe1, Hiroshi Fujii2 Institute for Molecular Science, Okazaki 444-8585, Japan 1 E-mail: [email protected] 2 E-mail: [email protected]

High-valent oxo metalloporphyrins have been known as reactive intermediates in the catalytic cycles of many heme enzymes and in the oxidation reactions mediated by synthetic metalloporphyrins. In this review, we survey studies of high-valent oxo metalloporphyrins in the past decade. In Sect. 2, electronic structures and magnetic properties of compound I and compound II species in various heme enzymes are discussed on the basis of recent studies of synthetic model complexes of the compound I and compound II. In Sect. 3, spectroscopic properties of oxo manganese porphyrin complexes and their application to biological systems are summarized. In the Sects. 4 and 5, synthesis and characterization of oxo chromium and oxo ruthenium porphyrin complexes are reviewed. In the last section, we have summarized the synthesis and characterization of Mo, Nb, Ti, and V oxo metalloporphyrins complexes. Keywords: Cytochrome P450, Peroxidase, Heme, Oxo, Metalloporphyrin

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

62

2

O@Fe Porphyrins and Related Complexes . . . . . . . . . . . . . . . .

63

2.1

2.2.5 2.3

Physical Properties and Characteristics of High Valent Intermediates in Heme Enzymes . . . . . . . . . . . . . . . . Studies on the Preparation and Characterization of Oxo-Ferryl Porphyrin Model Complexes Related to Cytochrome P-450 and Peroxidases . . . . . . . . . . . . A1u and A2u . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ligand Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chlorin Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . Isoelectronic Forms of Oxo-Iron(IV) Porphyrin p-Cation Radical . . . . . . . . . . . . . . . . . . . . . . . . . . . . Oxo-Iron(IV) Porphyrins . . . . . . . . . . . . . . . . . . . . . . Development of Heme Protein Engineering . . . . . . . . .

.......

63

. . . .

. . . .

65 66 69 70

....... ....... .......

71 71 72

3

O@Mn Porphyrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

75

3.1 3.2

Early Work on the O@Mn Porphyrins . . . . . . . . . . . . . . . . . . . Application of O@Mn Porphryin Chemistry to Biological Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Recent Advances in O@Mn Porphryin Chemistry . . . . . . . . . . .

75

2.2 2.2.1 2.2.2 2.2.3 2.2.4

3.3

. . . .

. . . .

. . . .

. . . .

. . . .

77 79

Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

62

Y. Watanabe á H. Fujii

4

O@Cr Porphyrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

80

5

O@Ru Porphyrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

81

6

Other O@M Porphyrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

82

7

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

84

1 Introduction The biological use of Fe@O structures for oxidative metabolisms has been intensively studied for over three decades. Readily available examples are heme enzymes including cytochrome P-450, peroxidases, oxidases, and so on [1±5]. Among those heme enzymes, cytochrome P-450 has received much attention due to its unusually high catalytic hydroxylation activity [2, 3, 6±8]. Cytochrome P-450cam is able to hydroxylate d-camphor to 5-exo-hydroxyd-camphor at a turnover number of >1000/min [9]. While, the active intermediate(s) of cytochrome P-450 responsible for the oxidations is still controversial, an oxo-ferryl porphyrin p-cation radical (O@Fe(IV) Por‡ ), equivalent to compound I of peroxidases, has been considered to be the active species [1±8]. Early studies mimicking cytochrome P-450-type reactions by metalloporphyrins were carried out by Tabushi and his coworkers in the late 1970s [10±12]. For example, they employed a Mn(III) hematoporphyrin/ NaOCl system for the oxidation of hydrocarbons, alcohols and ethers [10, 11]. Though Mn(IV) intermediates were proposed for the active species, no highvalent Mn complexes were characterized at the time. Efforts for the preparation of high-valent iron porphyrin complexes related to compounds I and II of peroxidases and catalase resulted in the preparation of O@FeIV(TPP) by electrochemical oxidation of [FeIII(TPP)]2O in CH2Cl2 in 1971 [13]. In 1979, Groves et al. reported the hydroxylation and epoxidation by PhIO in the presence of Fe(TPP) as a catalyst [14]. A few months later, Chang and Kuo were the ®rst to observe a transient intermediate similar to compound I of HRP by the reaction of Fe(III)(OEP) and excess ArAIO in CH2Cl2 at )45 °C [15]. Since then, many efforts have been devoted to the identi®cation and application of oxo-metalloporphyrins and related intermediates. In this review article, the authors have focused mainly on the formation and characterization of oxo-metalloporphyrin complexes both in biological and synthetic model systems. Thus, oxidations catalyzed by metalloporphyrin complexes are not described very much unless the reactions are highly related to mechanistic aspects of the high valent intermediate formation. For the convenience of readers who are interested in the metalloporphyrin catalyzed oxidations, we have listed the review articles in References [16±21].

Characterization of High-Valent Oxo-Metalloporphyrins

63

2 O@Fe Porphyrins and Related Complexes 2.1 Physical Properties and Characteristics of High Valent Intermediates in Heme Enzymes

High valent iron porphyrins have been detected in the catalytic cycles of peroxidases [22]. When peroxidase reacts with 1 equivalent of hydrogen peroxide, a metastable green intermediate named compound I is formed. The compound I oxidizes organic substrates by one electron to form the second red intermediate named compound II (Eq. 1). Resting State (ferric high spin) ‡ H2 O2 ! Compound I Compound I ‡ Substrate ! Compound II ‡ Oxidized Product Compound II ‡ Substrate ! Resting State ‡ Oxidized Product

…1†

The compound I of horseradish peroxidase(HRP) exhibits a unique absorption spectrum; a broad Soret band (400 nm) with loss of intensity and new broad peak at 651 nm [23]. The spectral feature of the compound I is similar to that of metalloporphyrin p-cation radical [24]. The absorption spectrum of the HRP compound II shows peaks at 418, 527 and 554 nm [23]. The MoÈssbauer spectra of the HRP compound I and II indicate that heme irons in compound I and II are in a ferryl oxidation state [25]. Based on these results, the electronic structures of compound I and II are proposed as oxo-iron(IV) porphyrin p-cation radical and oxo-iron(IV) porphyrin, respectively. Two types (a1u and a2u) of porphyrin p-cation radical states have been characterized from EPR and theoretical studies of metalloporphyrin p-cation radical complexes (Fig. 1) [26, 27]. The HRP compound I had been thought the a2u radical state from its absorption spectral feature [28]. However, recent studies have brought question the a1u/a2u assignment based on the absorption spectral feature and pointed out, rather, the a1u radical state for HRP compound I on the basis of several physicochemical measurements (Fig. 1)

Fig. 1. Electron spin distribution of porphyrin atomic orbitals with a1u (left) and a2u (right)

symmetries. Black and white circles represent signs of the upper lobe of the pp AOs

64

Y. Watanabe á H. Fujii

[29±31]. An EPR spectrum of HRP compound I exhibits a broad signal around g = 2 resulting from weak magnetic interaction between ferryl iron spins and porphyrin p-cation radical spin [32, 33]. Even in the porphyrin p-cation radical state, the HRP compound I exhibits well-resolved paramagnetic 1HNMR signals of prophyrin peripheral protons [34±36]. The porphyrin p-cation radical state of HRP compound I is also supported by 1H- and 14N-ENDOR spectra [37]. The presence of oxo ligands in HRP compound I is con®rmed by its 17O-ENDOR spectra [38]. The EXAFS measurements of the HRP compound Ê [39]. While the resonance I and II show the iron-oxygen distance to be 1.6 A Raman (RR) spectrum of HRP compound II shows the m(Fe@O) signal at 774 cm)1 [30, 40], the RR band of m(Fe@O) for HRP compound I has not been determined and still in controversy [30, 41±43]. The compounds I and II of cytochrome c peroxidase (CcP) are characterized by various spectroscopic methods. The absorption spectrum of CcP compound I is different from that of HRP compound I, but rather similar to that of HRP compound II [44, 45]. The EPR mesurement of CcP compound I exhibits the presence of a protein radical [46]. The MoÈssbauer spectrum of the CcP compound I is in good agreement with a ferryl oxidation state [47]. On the basis of these results, CcP compounds I and II are assigned as oxo-iron(IV) porphyrin with a protein radical and oxo-iron(IV) porphyrin, respectively. The ENDOR study of CcP compound I has identi®ed the primary site of the protein radical at tryptophan-191 [48]. The m(Fe@O) band is observed at 767 cm)1 for the compound I [49, 50]. The compound I of chloroperoxidase (CPO), which has a cystein thiolate axial ligand instead of the histdyl imidazole ligand in HRP, has absorption peaks at 367 and 689 nm [51]. Compared with HRP compound I, these peaks are blue and red-shifts, respectively. The EPR spectrum of the CPO compound I shows g = 1.64, 1.73, and 2.00, which indicate antiferromagnetic coupling between the ferryl iron and porphyrin p-cation radical spins [52]. This is in contrast to the weak magnetic interaction in HRP compound I. The RR spectrum of CPO compound I shows a m(Fe@O) vibration at 790 cm)1 [53]. In the catalytic cycle of catalase (CAT), the compound I is also an active intermediate which can disproportionate hydrogen peroxide to molecular oxygen and water (Eq. 2) [54]. Resting State (ferric high spin) ‡ H2 O2 ! Compound I Compound I ‡ H2 O2 ! Resting State ‡ H2 O ‡ O2

…2†

The compound I intermediate is not detected in the catalytic cycle because hydrogen peroxide serves as both oxidant and substrate. When catalase is oxidized by peracetic acid, a metastable compound I is observed [55]. The absorption spectrum of CAT compound I exhibits absorption at 400 and 662 nm, which is a typical spectrum of porphyrin p-cation radical complex [56]. Thus the CAT compound I is an oxo-iron(IV) porphyrin p-cation radical, as HRP compound I. Although the peak positions of CAT compound I are close to those of HRP compound I, the spectral feature of CAT compound I is obviously different from that of HRP compound I. This is interpreted as an

Characterization of High-Valent Oxo-Metalloporphyrins

65

effect of the axial tyrosyl ligand [57]. The EPR spectrum of compound I from Micrococcus luteus (MLC) showed EPR signals at g = 3.45 and 2.0, indicating weak ferromagnetic interaction of ferryl iron and porphyrin spins [58]. The EPR spectrum of CAT compound I from Proteus mirabilis (PMC) is also similar to that of MLC compound I at low pH, but changes to the signal at g = 2 like HRP compound I at neutral pH [59]. The RR band of m(Fe@O) has not been observed for CAT compound I, but observed at 775 cm)1 at neutral pH for CAT compound II from bovine liver [60]. From these spectral results, the CAT compound I is also assigned to the a1u radical state. Compound I and II species are also proposed in the catalytic cycle of cytochrome c oxidase (CcO), which is a terminal oxidase and catalyzes reduction of molecular oxygen with four electrons arriving through the respiratory chain [61, 62]. The high valent intermediates of CcO have been detected in ¯ow-¯ash experiments [63, 64]. Two intermediates having absorption peaks at 607 and 580 nm are reported and assigned as a peroxy species (compound I level) and ferryl species (compound II level), respectively [63]. The RR spectra of these intermediates assigned for m(O@Fe) show bands at 804 and 785 cm)1 [64]. 2.2 Studies on the Preparation and Characterization of Oxo-Ferryl Porphyrin Model Complexes Related to Cytochrome P-450 and Peroxidases

In 1981, Groves and coworkers found the reaction of FeIII(TMP)Cl (TMP: tetramesityl porphyrin) with m-chloroperbenzoic acid (mCPBA) in CH2Cl2-MeOH at )78 °C gave a green intermediate [65]. According to this color change, the b-pyrrole hydrogen resonance at 120 ppm in 1H NMR for FeIII(TMP)Cl was shifted to )27 ppm. The chemical shift of the intermediate obeyed the Curie law between )26 and )89 °C. The magnetic susceptibility determined by Evans method was 4.2 lB, indicating S = 3/2. The MoÈssbauer spectrum of the intermediate showed a quadrupole doublet centered at 0.05 mm/s (DEq = 1.49 mm/s). In addition, the UV-Vis spectrum of the intermediate showed a less intense Soret band with a broad band around 600 to 700 nm. All these physical properties support the green intermediate to be O@FeIVTMP‡ . The green complex shows EPR signals at g = 4.3, 3.9, and 1.99 at 4 K, indicating strong ferromagnetic coupling between ferryl iron (S = 1) and the porphyrin p-cation radical spin (S = 1/2) at J > +40 cm)1 [66]. This is in contrast to weak magnetic couplings observed for compounds I of peroxidases. Evidence for the O@Fe ligation was provided by EXAFS spectroscopic studies on O@FeIVTMP‡ [67]. For example, O@Fe distances for HRP compound I and O@FeIVTMP‡ were reported to be 1.61±1.64 and Ê , respectively [67]. Furthermore, Hashimoto et al. prepared 1.62±1.66 A 16 O@FeIVTMP‡ and 18O@FeIVTMP‡ in CH2Cl2/MeOH by using 16O- and 18 O-labeled mCPBA [68]. The resonance Raman (RR) bands for mO@Fe were observed at 828 and 792 cm)1 for 16O and 18O, respectively. In addition [54], Fe substituted samples gave the RR band at 832 cm)1. These isotope shifts are consistent with theoretical isotope frequency shifts expected for an O@Fe

66

Y. Watanabe á H. Fujii

oscillator. The O@Fe RR band is sensitive to the 6th ligand of O@FeIV Porphyrin‡ [69], thus, Kincaid et al. observed the O@Fe RR band of O@FeIVTMP‡ at 801 cm)1 in CH2Cl2 [70]. Detailed mechanistic studies on the reaction of FeIIITMP with mCPBA by Groves and Watanabe showed the transient formation of acylperoxo-FeIIITMP followed by the heterolytic OAO cleavage giving O@FeIVTMP‡ in polar solvent such as CH2Cl2 [71, 72]. Replacement of the solvent to less polar toluene or benzene changed the nature of OAO bond cleavage from the heterolysis to the homolysis (Scheme 1) [72, 73]. Substituent effects of porphyrin ring, peracid and 6th ligand on the homolysis and heterolysis were systematically examined and their relevance to cytochrome P-450 and peroxidases were discussed [74±75]. 2.2.1 A1u and A2u Change in the electron-negativity of meso-substituent alters the reactivity and electronic state of the oxo-iron(IV) porphyrin p-cation radical complex. Iron porphyrins having electron-withdrawing substituents such as tetra-2,6dchlorophenylporphyrin (TDCPP) and tetra-penta¯uorophenylporphyrin (TPFPP) (Fig. 2) are shown to be ef®cient catalysts of oxygen atom transfer

Scheme 1. The heterolysis and homolysis of acylperoxo-Fe(III) porphyrins

Fig. 2. The structures of meso-tetraarylporphyrins employed for synthesis of high valent oxo

porphyrin complexes

Characterization of High-Valent Oxo-Metalloporphyrins

67

reactions from iodosobenzene, hydrogen peroxide and so on [76]. O@FeIV(TDCPP‡ ) is prepared by the oxidation of FeIII(TDCPP)(ClO4) with ozone in acetonitile at )35 °C [77]. The EPR spectrum of O@FeIV(TDCPP‡ ) exhibits the signals at g = 2 in acetonitrile [78] and at g = 4.26, 3.62, 1.98 in dichloromethane-methanol (6:1) [66]. The EPR spectrum in dichloromethanemethanol is similar to that of O@FeIV(TMP‡ ) and interpreted by ferromagnetic interaction of the ferryl iron and the porphyrin p-cation radical with J = +38 cm)1 [66]. However, 1H-NMR spectrum of O@FeIV(TDCPP‡ ) shows much more up®eld shift of pyrrole proton ()66 ppm) than that ()27 ppm) of O@FeIV(TMP‡ ) [31], suggesting the change in the electronic state of the porphyrin radical. The comprehensive studies of the electronnegativity of meso-substituents revealed that the pyrrole proton signal shifts up®eld with increase in electron-negativity of meso-substituents; 2,4,6trichlorophenyl > 2,6-dichlorophenyl > 2-methyl-6-chlorophenyl > mesityl [31]. The spectral change is interpreted as the mixing of the a1u and a2u radical states by a vibronic coupling. Electron-withdrawing substituent at the mesopositions should lower the energy of the a2u orbital relative to the a1u orbital via the interaction of the aryl and porphyrin p-orbitals, since electron density is concentrated at the meso-position for the a2u orbital while the a1u orbital has a node (Fig. 1). This leads to a decrease in the a2u-a1u energy separation (Fig. 4) and stronger mixing of the a1u radical state to the a2u state would be expected with increase in the electron-negativity of meso-substituent. EPR spectra of these electron-negative p-cation radical complexes are similar to that of O@FeIV(TMP‡ ), indicating ferromagnetic coupling between the ferryl iron and the porphyrin radical spins [79]. When the meso-substituents are much more electron-negative such as penta¯uorophenyl, HOMO of the porpyhrin orbital is switched to the a1u orbital from the original a2u orbital. Thus, O@FeIV(TPFPP‡ ) shows large up-shift ()96 ppm) of pyrrole proton signal, consistent with the a1u radical state [80]. The EPR spectrum of O@FeIV(TPFPP‡ ) exhibits a broad signal around g = 2 as observed for HRP compound I suggests a weak interaction of the ferryl iron and the porphyrin p-cation radical spins [80]. This is also reasonable to the a1u radical state. The spin interaction of the a1u radical with the iron spins would be weak because of no spin density at the pyrrole nitrogen atoms in the a1u orbital, but that of the a2u would be strong because of large spin density at the pyrrole nitrogen atoms in the a2u orbital. Drastic change in the magnetic interaction with the change from the a2u radical to the a1u radical is also observed in the copper(II) porphyrin p-cation radical complexes [81]. The position of porphyrin substituents is also important to determine the electronic state of oxo-iron(IV) porphyrin p-cation radical (Fig. 3). While heme enzymes utilize protoporphyrin or its modi®ed forms as the prosthetic group, i.e., pyrrole b-substituted porphyrin, most of model complexes which provide oxo-iron(IV) porphyrin p-cation radical complexes are mesotetraphenylporphyrin derivatives because of their easy preparations. An oxoiron(IV) porphyrin p-cation radical complex bearing pyrrole b-substituents has been prepared and characterized by Fujii et al. using 2,7,12,17-tetramethyl3,8,13,18-tetramesityporphyrin (TMTMP) [82]. By changing the substituents

68

Y. Watanabe á H. Fujii

Fig. 3. The structures of pyrrole b-substituted porphyrin complexes

from the meso-positions to the pyrrole b-positions, spectroscopic properties are altered to those of compounds I of heme enzymes. The absorption spectrum of O@FeIV(TMTMP‡ ) is similar to that of HRP compound I, but very different from that of O@FeIV(TMP‡ ) [31, 82]. The 1H-NMR spectrum of O@FeIVTMTMP‡ exhibits large downward shift of the pyrrole methyl proton at 140 ppm and small shift of the meso-proton at 53 ppm at )70 °C, consistent with the spin distribution of the a1u orbital [31, 82]. The EPR spectrum of O@FeIVTMTMP‡ shows signals at g = 3.1 and 2.0, which is close to that of compounds I of Micrococcus luteus catalase and ascorbate peroxidase [79]. The EPR spectrum is interpreted by a weak ferromagnetic interaction between ferryl iron and porphyrin radical spins, as expected for the a1u radical state. The Raman spectrum of O@FeIVTMTMP‡ (ClO4) shows the m(Fe@O) band at 801 cm)1 [83], which is the same position as that of O@FeIVTMP‡ (ClO4) [84]. This means that the a1u/a2u radical state does not change the Fe@O bond strength. In contrast to the meso-substituted porphyrin, the electronic structure is not altered by the electron-withdrawing substituents at pyrrole b-positions (Fig. 4) [31]. This can be also explained by the spin distributions

Fig. 4. Change in the a1u and a2u orbitals with increasing in the electron-negativity of the

substituent

Characterization of High-Valent Oxo-Metalloporphyrins

69

of the a1u and a2u orbitals (Fig. 1). Since the a1u and a2u orbitals have small spin densities at the pyrrole b-positions, both orbitals are slightly stabilized by electron-withdrawing substituents, as shown in Fig. 4. Thus, pyrrole bsubstituents do not drastically alter the orbital occupancy of the pyrrole bsubstituted oxo iron(IV) porphyrin p-cation radical, consistent with the a1u radical state. On the basis of these results, the a1u radical state of HRP compound I is proposed [31]. 2.2.2 Ligand Effects The effect of an axial ligand on the reactivities of oxo-iron(IV) porphyrin p-cation radicals has been studied in connection with various axial ligands in heme enzymes. The epoxidation reactivity of oxo-iron(IV) porphyrin p-cation radicals was found to be drastically altered by its axial ligand. O@FeIV(TMP‡ )(F) rapidly reacted with styrene to form styrene oxide, though O@FeIV(TMP‡ )(X), in which X = methanol, chloride and acetate, were less reactive. In addition, O@FeIV(TMP‡ )(ClO4) did not react with ole®ns [85]. The RR m(Fe@O) band is affected by the nature of the trans-axial ligand, for example, the RR m(Fe@O) band was observed near 835 cm)1 for perchlorate and trifrate complexes while near 800 cm)1 for the m-chlorobenzoate, ¯uoride and chloride complexes [84]. In these studies, alcohols, halides and some other weakly coordinating anions have been examined. The RR spectrum of O@FeIV(TMP‡ )(ImH) showed m(Fe@O) vibration at 810 cm)1 and that of O@FeIV(TMP)(ImH) appared at 814 cm)1, the Raman band is just 4 cm)1 down-shifted by porphyrin ring oxidation [86]. Very recently, oxo-iron(IV) TMTMP p-cation radical complexes bearing imidazole and p-nitrophenol have been prepared as models for the compounds I of peroxidase and catalase (Scheme 2) and characterized by absorption and 1H-NMR spectroscopies [87]. The absorption spectra of O@FeIVTMTMP‡ (ImH) and O@FeIVTMTMP‡ (p-NO2-PhO) were quite close to those of compounds I of peroxidase and catalase, respectively [87]. 1H-NMR

Scheme 2. Introduction of biologically relevant trans-axial ligand into a oxo-iron(IV)

porphyrin p-cation radical complex

70

Y. Watanabe á H. Fujii

of O@FeIVTMTMP‡ (ImH) exhibits paramagnetic shifts of the pyrrole bmethyl protons similar to those of HRP compound I. O@FeIVTMTMP‡ (ImH) is readily assigned as the a1u radical state even with coordinating imidazole as an axial ligand. The preparation of high valent iron porphyrin bearing thiolate ligand as a model for cytochrome P-450 has been hampered by high reactivity of the thiolate ligand with air and oxidants [88, 89]. Recently, air-stable iron(III) porphyrin complexes having thiolate ligands have been synthesized by connecting the thiolate ligands with porphyrin through a covalent bond and/or by surrounding the thiolate ligand with pivaroyl groups (Fig. 5) [90± 92]. Because of the strong push-effect from the trans-axial thiolate ligand, the oxygenation activities of these thiolate complexes are much higher than those of non-thiolate complexes. Very recently, an intermediate bearing absorption peak at 388 nm has been observed and tentatively assigned as oxo iron(IV) porphyrin p-cation radical state [93]. 2.2.3 Chlorin Derivatives The effect of porphyrin ring reduction (i.e., chlorin) is studied as a model for the compounds I of chlorin-containing heme enzymes such as Neurospora crassa catalase [94]. Oxo-iron(IV) tetramesitylchlorin (TMC) p-cation radical, O@FeIV(TMC‡ ) (Fig. 6) is prepared by mCPBA oxidation in dichloromethane at )80 °C and characterized by absorption and 1H- and 2D-NMR spectroscopies [95]. The 1H-NMR shifts of pyrrole and mesityl protons of O@FeIV(TMC‡ ) are consistent with an a2 radical character (a2 symmetry corresponds to the a1u in the D4h symmetry). This is in contrast to the a2u state assigned to O@FeIV(TMP‡ ). The porphyrin ring reduction destabilizes the energy of the a2 orbital relative to that of the b2 orbital (the a2u in D4h symmetry). The RR band of O@FeIV(dihydroxy-dimethyl-TMC‡ ) at 832 cm)1 is shifted to 800 cm)1 on oxidation with 18O-mCPBA. Thus, the m(Fe@O) bond is very close to that of O@FeIV(TMP‡ ) [96]. O@FeIV(dihydroxy-dimethylTMC‡ ) exhibits EPR signals at g = 4.23, 3.80, and 2.02 arisen from the S = 3/2 state [96]. The MoÈssbauer parameters of O@FeIV(dihydroxy-dimethyl-TMC‡ )

Fig. 5. Air-stable iron(III) porphyrin complexes having thiolate ligands

Characterization of High-Valent Oxo-Metalloporphyrins

71

Fig. 6. Fe tetramesitylchlorin, Fe(TMC)

are d = 0.01 mm/s and DEq = +1.25 mm/s, which is almost identical with those of O@FeIV(TMP‡ ) [96]. Interestingly, the oxidation of norborene by O@FeIV(TMC‡ ) proceeds much slower than that by the corresponding TMP complex [95]. 2.2.4 Isoelectronic Forms of Oxo-Iron(IV) Porphyrin p-Cation Radical Iron(V) porphyrin, which is an isoelectronic form of the oxo-iron(IV) porphyrin p-cation radical, has been prepared by introducing methanol into an oxo-iron(IV) TDCPP p-cation radical as an axial ligand at )90 °C [97, 98]. The iron(V) oxidation state can be prepared by the introduction of strong electron-withdrawing substituents on the porphyrin ring and by the coordination of methoxide. The absorption spectrum of O@FeV(TDCPP) is similar to that of O@FeIV(TDCPP). The iron(V) oxidation state of the complex is con®rmed by iodometric titration. Oxo-iron(V) porphyrins are reactive toward ole®n and O@FeV(TDCPP) oxidized norborene to norborene oxide in 50% yield at )90 °C. The other isoelectronic forms of oxo-iron(IV) porphyrin p-cation radical are iron(III) porphyrin N-oxide [99±101] and iron(III) porphyrin dication complexes (Fig. 7) [102]. These complexes have been prepared and characterized by absorption, 1H-NMR, EPR and Raman spectroscopies. 2.2.5 Oxo-Iron(IV) Porphyrins Through the works on the autooxidation of Fe(II) porphyrin complexes, Balch et al. showed the transient formation of Fe(III)AOAOAFe(III) species [103].

Fig. 7. Fe porphyrin N-oxide and Fe porphyrin dication

72

Y. Watanabe á H. Fujii

More importantly, they found that the introduction of N-methylimidazole (MeIm) to the Fe(III)AOAOAFe(III) complex affords O@FeIV(Por)(MeIm) (Eq. 3) [104, 105]. O@FeIV(Por)(MeIm) is able to oxidize triphenylphosphine even at )80 °C in toluene over a period of several hours to give triphenylposphine oxide quantitatively [105]. …Por†FeIII AOAOAFeIII …Por† ‡ 2 Melm ! 2 O ˆ FeIV …Por†…Melm†

…3†

Very recently, a series of independent papers have appeared concerning the conversion of Fe(III) porphyrins to the corresponding O@Fe(IV) complexes. Groves et al. prepared O@FeIV(TMP) from FeIII(TMP‡ )(ClO4)2 by passing through basic alumina [106]. Though O@Fe(IV) porphyrins are believed to be poor oxidants for ole®n oxidation, they found the epoxidation of substituted styrenes by isolated O@FeIV(TMP). Kinetic data and product analyses in benzene indicated that the ®rst and rate determining step is attack of the ferryl oxygen of O@FeIV(TMP) on the ole®n p-bond to give a carbon radical intermediate, following by a fast reaction with O2 to afford benzaldehyde or epoxides with loss of the con®guration of the starting ole®n. These mechanistic features are quite different from the results by O@FeIV(TMP‡ ) (Scheme 3). The ®rst electrochemical oxidation of haloiron(III) porphyrins generally affords porphyrin centered p-cation radicals [107], while the same oxidation of hydroxoiron(III) porphyrins gives oxoiron(IV) porphyrins [108±111]. The E1/2 values of eight FeIII(Por)(OH) was plotted against the E1/2 values of the corresponding FeIII(Por)(Cl) [106]. The correlation was found to be unity, indicating that ®rst oxidation of FeIII(Por)(OH) proceed through FeIII(Por‡ )(OH) to afford O@FeIV(Por). A similar conclusion was also obtained by Gray et al. upon photoinduced oxidation of microperoxidase-8, the heme octapeptide derived from enzymatic cleavage of cytochrome c [112]. Very recently, Hamachi et al. have succeeded in detecting a transiently formed FeIII(Por‡ ) as a precursor of O@Fe(IV) by using semisynthetic myoglobins having covalently appended Ru(byp)3 [113]. 2.3 Development of Heme Protein Engineering

Peroxidases and catalase react with hydrogen peroxide to afford oxo-ferryl porphyrin p-cation radicals so called compound I through the heterolysis of an

Scheme 3. The oxidation of ole®ns by O@Fe(IV) porphyrins

Characterization of High-Valent Oxo-Metalloporphyrins

73

iron bound hydroperoxide [4, 5]. Homolytic OAO bond cleavage is usually observed for the reactions of hydroperoxide with many transition metal ions to form hydroxyl radical [114] , thus, general acid-base function of the distal histidine in those enzymes is believed to be crucial for avoiding radical process (Scheme 4) [115]. Quantitative roles of the distal residues of cytochrome c peroxidase were examined by replacing these residues with other amino acids by site-directed mutagenesis. For example, mutation of the distal histidine to leucine was found to depress the compound I formation rate from 3 ´ 10)7 to 2 ´ 10)2 M)1 s)1 [116, 117]. In the case of cytochrome P-450, the actual intermediate responsible for the oxygenation is still controversial, however, Ishimura et al. and Sligar et al. observed the important participation of the distal threonine 252 residue in cytochrome P-450cam as a proton donor for the heterolytic OAO bond cleavage as discussed in Scheme 4 [118, 119]. Recent mechanistic studies have shown the presence of a proton network in the distal site of P-450cam to allow the activation of a peroxo intermediate [120]. Very recently, a crystal structure of nitric oxide synthase-oxygenase dimer with pterin and the substrate (L -arginine) has been reported by Crane et al. [121]. Even in this case, L -arginine stays very close to the heme iron and is capable of serving as a proton donor (Scheme 5). These protons could work together with cysteinate ligation for the compound I formation in cytochrome P-450. By comparison of the distal site structures of Mb [122] and cytochrome c peroxidase [123], two major differences are apparent; 1) the location of two histidines and 2) the histidine of CcP makes a hydrogen bond with asparagine (Asn 82) while the histidine of Mb stands alone (Fig. 8) [122]. According to the hydrogen bond, the distal histidine in CcP is much more basic than that of Mb [122]. As shown in Scheme 4, the proton abstract from hydrogen peroxide is the initial step of compound I formation, suggesting that the basicity of the distal histidine could play an important role. In fact, replacement of asparagine 70 in HRP, which interacts with the distal histidine (His 42) of HRP through the hydrogen bond, with valine decreased the compound I formation less than 10% of the native HRP [123].

Scheme 4. Roles of the distal histidine for the formation of compound I

74

Y. Watanabe á H. Fujii

Scheme 5. Proposed proton assisted formation of compound I in the ®rst step of NO

synthesis by NOS

Fig. 8. Superimposition of the distal histidines of CcP and Mb

Characterization of High-Valent Oxo-Metalloporphyrins

75

The location of the distal histidine in the peroxidase should also be very important for the compound I formation. According to the crystal structures of CcP and beef liver catalase [124], the distances between the heme iron and Ê , respectively [123, 124]. About 5.0± Ne of the distal histidine are 5.6 and 4.8 A Ê 5.5 A could be the distance suitable for the hydrogen bond between the protonated distal histidine and the terminal oxygen of the iron bound hydroperoxide (Fig. 9) [125]. On the other hand, the distance between the Ê [122] and it heme iron and Ne of the distal histidine (His43) in Mb is 4.3 A seems too close to make a peroxidase type hydrogen bond; instead, the protonated distal histidine readily makes the hydrogen bonds with both oxygens of the hydroperoxide (Fig. 9) [126]. To investigate this hypothesis, Watanabe et al. have prepared F43H/H64L Mb, in which the distal His64 was moved to the position of Phe43 [127]. On the basis of the mutant crystal structure, the distance of Ne of the His43 and the Ê [128]. F43H/H64L Mb oxidized heme iron was determined to be about 5.5 A thioanisole and styrene 200±300 times faster than the wild-type Mb with very high enantio-selectivity [127a, 129]. For example, thioanisole and transb-methylstyrene were oxidized to R-sulfoxide and 1R, 2R-epoxide in e.e. of 85% and 96%, respectively [129]. Furthermore, the reaction of F43H/H64L Mb and mCPBA gave a transient intermediate characterized as compound I, which gradually photo-reduced to compound II (Fig. 10) [127b, 128]. Finally, compound I of F43H/H64L Mb has been shown to react with substrates including H2O2 (i.e. catalase activity) by stopped ¯ow measurements [127b,c, 129, 130].

3 O@Mn Porphyrins 3.1 Early Work on the O@Mn Porphyrins

Hill and Groves independently reported the oxidation of hydrocarbons by PhIO in the presence of MnIII(TPP)Cl [131, 132]. As shown in Scheme 6, the

Fig. 9. Schematic drawing of two types of hydrogen bond by the distal histidines of CcP and Mb

76

Y. Watanabe á H. Fujii

Fig. 10. Time dependent spectral changes of F43H/H64L Mb upon the addition of mCPBA

reaction is expected to proceed through the radical mechanism on the basis of the products [131, 132]. Upon the addition of PhIO to a MnIII(TPP)Cl solution, a reactive intermediate tentatively assigned as O@MnVTPP was observed. Through crystallographic studies, Hill et al. demonstrated a series of Mn(IV) structures (Fig. 11) [133±136], whose UV-Vis spectra are very similar to those observed in earlier work. For the MnIV(TPP)(OMe)2 complex, bond distances Ê ; two MnAN1 bonds, 1.993(3) A Ê; are as follows: two MnAO bonds, 1.839(2) A Ê Ê two MnAN2 bonds, 2.031(3) A; two OAC bonds, 1.387(3) A [134]. The magnetic susceptibility of MnIV(TPP)(OMe)2 determined in solution and the solid state at 25 °C gave leff = 3.9 lB, consistent with a high-spin d3 con®guration of Mn(IV). Meunier et al. also isolated Mn(IV) complexes by using NaOCl(aq.) as an oxidant [137]. According to X-ray absorption spectroscopies (EXAFS, XANES), one axial ligand was well de®ned Ê ) and assigned to a MnAO single bond and the other (R = 1.84 ‹ 0.02 A Ê ) and assigned to a loosely bound water bond was much larger (R = ca. 2.3 A molecule [138]. EPR spectra of those Mn(IV) complexes showed strong g?  4:0 and weak g?  2:0 signals characteristic of a high-spin d3 con®guration [139]. Mechanistic studies on the formation of high-valent oxo-Mn porphyrin complexes were carried out by Groves and Watanabe [140]. A peroxoMnIII(TMP) was found to react with acylating reagents to afford an acylperoxo-MnIII(TMP) in CH3CN at low temperatures. By employing mCPBA, they also observed the acylperoxo-Mn(III) adduct formation in CH2Cl2, which

Scheme 6. Mn porphyrin catalyzed oxidation of hydrocarbons

Characterization of High-Valent Oxo-Metalloporphyrins

77

Fig. 11. Mn(IV) porphyrin complexes well-characterized by X-ray crystallography

readily decomposed to afford O@Mn(V) porphyrins [140]. A similar oxygen activation of peroxo-Mn(III) complexes by CO2 were also reported [139b]. The O@MnVTMP is EPR silent immediate after the formation, however, EPR signals assignable to O@MnIV(TMP) was observed in a few seconds even )78 °C in CH2Cl2, implying O@MnV(TMP) is very unstable even at low temperatures [141]. Substituent effect of the aromatic ring of peroxybenzoic acid on the formation of O@MnV(TMP) indicated heterolytic cleavage of the OAO bond in the acylperoxo-Mn adducts proceed under acidic conditions, while homolysis is a favorable process in the presence of hydroxide anion [141]. The very similar homolysis/heterolysis nature of acylperoxo-Fe(III) porphyrin has been also demonstrated by Groves and Watanabe [71±73]. Schappacher and Weiss also reported both homolytic and heterolytic OAO bond cleavage for MnIII(Por)(OOACO2) [139b]. 3.2 Application of O@Mn Porphryin Chemistry to Biological Systems

The heme prosthetic group of heme proteins has been replaced by many manganese(III) porphyrins [142±148]. UV-Vis spectra of Mn(III) substituted heme proteins are usually very similar to those of Mn(III) porphyrins in aqueous solutions. Anionic ligands such as N3 , F) and CN) does not bind to Mn reconstituted myoglobin (Mb) and horseradish peroxidase (HRP) as evidenced by no UV-Vis spectral change upon the addition of these external ligands, while Mn substituted cytochrome c peroxidase (CcP) gives altered

78

Y. Watanabe á H. Fujii

spectra [142]. Modi, Mitra and co-workers examined the binding of SCN) and CN) to Mn(III) HRP by relaxation rate measurements of 15N resonance [147a]. Though the addition of CN) to Mn(III) HRP did not alter its UV-Vis spectrum, 15N-T1 studies indicated that the interaction of CN) with Ê from Mn(III) HRP is very weak and away from the metal center, ca. 6.8 A the Mn(III) center. Non-binding of anionic ligands may be caused by strong binding of a hydroxide ion as the distal site ligand. Yonetani and Asakura systematically prepared Mn(III) substituted CcP, HRP and sperm whale Mb with proto-, hemato-, meso- and deuteroporphyrins [144]. Mn(III) peroxidases reacted with H2O2 to form intermediates whose UV-Vis spectra were similar to those of well characterized Mn(IV) porphyrins described above, while these Mn peroxidases showed very low peroxidase activities. The intermediates are able to oxidize peroxidase substrates and retain 2 oxidizing equivalents per mole of the enzymes. On the other hand, Mn(III) Mb neither afforded the intermediate nor exhibited peroxidase activity by the reaction with H2O2. Upon the addition of H2O2, Mn(III) CcP gave an EPR signal at g = 2, however, the spin concentration of the signal was estimated to be less than 5% of the molar concentration of the enzyme. Thus, Yonetai et al. concluded that the free radical was not the major component of the intermediate [144]. Later, the axial ligand in the Mn(IV) HRP prepared by H2O2 was postulated to be a hydroxo ligand by Makino et al. [148]. Groves, Spiro and coworkers further studied the reaction of Mn(III) HRP with mCPBA instead of H2O2 [146]. The intermediate, they called Mn HRP-I, showed absorption maximum at 412 nm, typical for Mn(IV) porphyrins, and stable enough to isolate by Sephadex chromatography at 4 °C. The EPR spectrum of the isolated Mn HRP-I exhibited signals at g 5.25 and 2.08, characteristic of Mn(IV) and a free radical. Titration of Mn HRP-I to MnIIIHRP required 2 equivalent of thiosalicylate. The ®rst equivalent exclusively eliminated the Mn(IV) signal. The second equivalent titrated the g 2 signal. At the same time, 413.1 nm excited RR spectrum of Mn HRP-I showed a band at 622 cm)1, which shifted to 592 cm)1 by the mCPBA-18O preparation. Likewise synthetic O@Mn(IV) porphyrins, 622 cm)1 band was also observed even in H2 18 O, consistent with slow oxo exchange with solvent H2O observed for model systems (vide infra). Further, the 622 cm)1 band was down-shifted by 4 cm)1 in D2O [146]. Gray et al. have prepared manganese(III) substituted microperoxidase-8 (Mn MP8) and examined its reactions with H2O2 and Ru(byp)3‡ [149]. A 3 transient intermediate formed by H2O2 reaction exhibited the Soret at 401 nm and was less stable and readily reduced to MnIII MP8 in seconds. The same product was also observed by the photoinduced oxidation of MnIII MP8 in the presence of Ru(byp)2‡ 3 . On the basis of the spectroscopic characteristics, Gray et al. proposed the intermediate to be O@MnIV MP8. More importantly, the low reactivity of Mn(III) substituted heme proteins with peroxides was attributed to reduced ``push effect'' from the other axial ligand [149].

Characterization of High-Valent Oxo-Metalloporphyrins

79

3.3 Recent Advances in O@Mn Porphryin Chemistry

As mentioned above, isolated high valent Mn porphyrin complexes never had the O@Mn bond in early work. Finally, O@MnIV(TMP) and O@MnIV(TMP)(OH) were isolated and characterized [150]. The MnIV@O stretching frequency for the former complex was identi®ed at 754 cm)1 (18O: 722 cm)1) by both RR and IR spectroscopies. On the other hand, 6-coordinate O@MnIV(TMP)(OH) showed the O@Mn stretching frequency at 712 cm)1 while FeAOH was not apparent. More than 100 cm)1 lowering of O@Mn band in HRP was attributed to the imidazolate character of the proximal histidine ligand in HRP. Likewise the O@Fe porphyrins, O@Mn(V) porphyrins and O@Mn(IV) porphyrins exhibit very different reaction pro®les as summarized in Scheme 7. Ayougou et al. prepared a Mn(IV)(TpivPP) complex by KO2/CO2 at )70 °C according to Scheme 8 [139]. The sample was applied to X-ray absorption spectroscopy (XAS). The results were consistent with the formation of a ®vecoordinate O@Mn(IV) complex, in which the MnAO bond distance was 1.69 ‹ 0.03. Due to the instability of O@Mn(V) porphyrins, their formation is usually demonstrated by oxidation of organic substrates such as ole®ns, alcohols, and

Scheme 7. Comparison of O@Mn(V) and O@Mn(IV) porphyrin catalyzed epoxidation

Scheme 8. Preparation of O@MnIV(TpivPP)

80

Y. Watanabe á H. Fujii

hydrocarbons [151±153]. Stereospeci®c epoxidation shown in Scheme 7 is also an example of characteristic nature of O@Mn(V) porphyrins [150a]. Thus, O@MnV(TMP) was only observed in CH2Cl2 as a transient intermediate and reduced in a few seconds to O@Mn(IV) species even at )78 °C [141]. Very recently, Groves et al. successfully prepared an O@Mn(V) porphyrin complex by employing water soluble tetra-(N-methylpyridyl)porphyrin (TMPyP) at room temperature [154]. The reaction of MnIII(TMPyP) with mCPBA, HSO5 , or ClO- has afforded a short lived intermediate with the Soret maximum at 443 nm with second order kinetics, ®rst order in [MnIII(TMPyP)] and [oxidant]. The intermediate decays to O@MnIV(TMPyP) as a relatively stable species (kmax: 428 nm). The O@Mn(V) intermediate is highly reactive toward ole®ns while O@Mn(IV) complex is not capable of the reactions under the conditions. In addition, the oxo oxygen was demonstrated to be exchangeable with H2 18 O as described in Scheme 7. Unlike the O@Mn(IV) porphyrins, neither crystal structures nor vibration of O@Mn(V) porphyrins are available. However, Collins et al. have reported crystal structures of oxo-Mn(V) complexes having non-porphyrin ligands (Fig. 12) [155]. The oxo-Mn(V) complexes are diamagnetic and the OAMn Ê . According to the short OAMn bond bond distances are 1.548±1.555(4) A distance, they have attributed the oxo-Mn(V) bonds to be a triple bond formulation [156]. Thus, OAMn stretching bands appears around 970± 979 cm)1 (933±942 cm)1 for 18O), depending on the structure of the ligand. In the case of the oxo-Mn(V) complex having a macrocyclic tetraamide ligand, Ê above the mean plane. Likewise the O@Mn(V) the Mn atom sits 0.60 A porphyrins, the oxo ligand exchanges in 18O-labeled water [157].

4 O@Cr Porphyrins The methodology used for the O@Mn and O@Fe porphyrin preparation is applicable for the O@Cr(V) porphyrin synthesis. CrIII(TPP)Cl catalyzes hydroxylation and epoxidation with PhIO in good yields [158]. Treatment of CrIII(TPP)Cl either with PhIO or mCPBA gives a red intermediate assigned as O@CrV(TPP). A dilute solution of O@CrV(TPP) (ca. 10)5M) is stable for several hours at room temperature, addition of ole®ns or alcohols causes rapid regeneration of CrIII(TPP)Cl. Clear evidence for the formation of O@Cr was provided by IR measurements; m16 O@Cr = 1026 cm)1, m18 O@Cr = 982 cm)1.

Fig. 12. Non-porphyrin Mn complexes, which are able to form O@Mn(V) complexes

Characterization of High-Valent Oxo-Metalloporphyrins

81

The O@Cr(V) was also demonstrated by the EPR spectra of O@CrV(TPP), in which 53Cr and 17O hyper®ne splitting was evident [158, 159]. In 1981 and 1982, three groups independently reported the preparation of O@CrIV porphyrin complexes as diamagnetic species [160, 162]. According to the crystal structures of O@CrIV(TPP), O@Cr was determined to be 1.62(2) by Budge et al. [160] and 1.572(6) by Groves et al. [161] respectively. The mO@Cr was observed at 1025 cm)1 (981 cm)1 for 18O), very similar to that of O@CrV(TPP). A less common preparation of O@Cr(IV) complexes was also reported by Yamaji et al., i.e., the photochemical formation of O@CrIV(TPP) from ONOACrIII(TPP) [163]. O@CrIV porphyrin complexes are less reactive than the corresponding O@CrV complexes, though quantitative conversion of benzyl alcohol to benzaldehyde and triphenyl phosphine to its oxide were reported. Greager and Murray prepared O@CrV(TPP) by electrochemical oxidation of O@CrIV(TPP) [164]. Likewise CrIII(Por), CrIII(salen) complexes react with PhIO to yield O@CrV(salen)+X) complexes, which oxidize ole®ns in reasonable yields [165]. The mO@Cr band appeared in the range of 935±997 cm)1 depending on the nature of a counter anion (X)).

5 O@Ru Porphyrins Because of the stable formation of high oxidation states, a wide variety of O@Ru complexes have been prepared. Most of those complexes are prepared by using non-porphyrin ligands [166±169]. For the ruthenium porphyrin complexes, l-oxo-bis-ruthenium(IV) complexes, [(RuIV(Por)(OR)]2O, were prepared and characterized by X-ray crystallography [170, 171]. Leung et al. oxidized RuIII(OEP)(PPh3)2 with PhIO in CH2Cl2 to afford a green intermediate, which was tentatively assigned as O@RuIV(OEP‡ )Br [172]. A stoichiometric titration with PPh3 gave O@PPh3 and [RuIV(OEP)2O] quantitatively. By using RuIII(OEP)(PPh3)2 as a catalyst, PhIO was able to oxidize hydrocarbons such as norbornene, styrene, cyclohexene and cyclohexane. Groves and Quinn prepared highly oxidized trans-bis-oxo-ruthenium porphryin, (O@)2RuVI(TMP) by the reaction of RuII(TMP)(CO) and mCPBA in CH2Cl2 at room temperature [173]. Further development of (O@)2RuVI(TMP) preparation by aerobic oxidtion of RuII(TMP) according to the mechanism shown in Scheme 9 allowed to proceed aerobic epoxidation [174]. Homogeneous catalytic oxidations using ruthenium porphyrin complexes are summarized by James [175].

Scheme 9. Aerobic epoxidation catalyzed by Ru(TMP)

82

Y. Watanabe á H. Fujii

As expected, (O@)2RuVI(TMP) complex is diamagnetic and exhibits wellresolved 1H-NMR signals in diamagnetic region [173]. The complex can be isolated as a solid form at room temperature [176], but the X-ray crystal structure has not been reported. The crystal structure of analogous dioxoosÊ for Os@O bond [177]. The mium(VI) porphyrin has been determined; 1.743 A infrared spectrum of (O@)2RuVI(TMP) shows a strong band at 821 cm)1 which is assigned as m(Ru@O) band. The m(Os@O) band for (O@)2OsIV(OEP) is assigned as 825 cm)1. (O@)2RuVI(OEP) and (O@)2RuVI(TPP) which are not sterically bulky porphyrin complexes are prepared by the oxidations of CORuII(OEP) and CORuII(TPP) with mCPBA in the presence of alcohol (methanol or ethanol), which occupies the vacant coordination site of a ruthenium(IV) intermediate [176]. Later, the ruthenium(IV) intermediate is identi®ed as dihydroxy ruthenium(IV) porphyrin complex, but not an oxo ruthenium(IV) porphyrin [178, 179]. The cyclic voltammogram of (O@)2RuVI(TMP) in dichloromethane shows one-electron reversible oxidation couples around 0.6 V vs SCE [173]. The oxidized product was characterized dioxo ruthenium(VI) porphyrin p-cation radical complex [173, 180] O@RuIV(TMP) complex has been characterized both in the synthesis of (O@)2RuVI(TMP) from CORuII(TMP) and in the reduction of (O@)2RuVI(TMP) with triphenylphosphine [181]. O@RuIV(TMP) is unstable toward disproportionation which affords (O@)2RuVI(TMP) and RuII(TMP) [181]. While (O@)2RuVI(TMP) oxidizes an ole®n to an epoxide, O@RuIV(TMP) does not oxidize the ole®n but reacts with triphenylphosphine [180]. Recently, highly ef®cient oxidizing systems using ruthenium porphyrin have been reported [182]. The systems employ 2,6-dichloropyridine N-oxide as the oxidant and oxidize adamantane to adamantan-1-ol (up to 68%), adamantan-1,3-diol (25%) and adamantan-2-one (1%). Tentative formulation of an active intermediate includes oxo-Ru(V) porphyrin, oxo-Ru(IV) porphyrin p-cation radical or oxo-Ru species haiving the pyridine N-oxide ligand.

6 Other O@M Porphyrins In connection with the intermediates of native heme enzymes introduced in the previous section, various oxo metalloporphyrin complexes have been synthesized. Oxo metalloporphyrin complex shows various reactivities depending on the central metal ion and its oxidation state. For example, with oxophilic metals such as titanium and vanadium, oxo metalloporphyrins are inert and unable to transfer an oxygen atom to organic substrates at room temperature. On the other hand, oxo metalloporphyrins of iron, manganese, and chromium are known to be capable of oxidizing organic substrates. This section is focused on synthesis and characterization of various high valent metal oxo porphyrin complexes and also discusses their relation to the nature of active iron-oxo species in heme enzymes. Metalloporphyrin complexes with groups IVB to VIIB readily binds oxygen atom(s) to form both mononuclear and binuclear complexes. Very early

Characterization of High-Valent Oxo-Metalloporphyrins

83

studies by Buchler et al. showed a series of high valent oxo and l-oxo complexes of Hf(V), Mo(V), Nb(V), W(V) and Re(V) [183]. A crystal structure of O@MoVAOAMoV@O showed the O@Mo and MoAO distances to be Ê , respectively [184]. Replacement of the O@Mo(V) unit 1.708(3) and 1.936(3) A V in the O@Mo AOAMoV@O to ClAMo(IV) caused the OAMo distance in MoIVAOAMoIV to be 1.851(6) [185]. On the other hand, O@MoIV(TPP) Ê [186]. O@MoIV(TPP) reacts showed the O@Mo bond distance being 1.656(6) A V V with O2 to form O@Mo AOAMo @O. When sterically hindered porphyrin such as TMP is used instead of TPP, an oxygen adduct of molybdenum(VI) complex, O@MoVI(TMP)(O2), is characterized from the reaction of O@MoIV(TMP) with O2 [187]. The O@Mo distances for cis-dioxo-MoVI(TPP) Ê with a saddle shape porphyrin were also reported to be 1.709(9) and 1.744(9) A core (Fig. 13a) [188]. Interestingly, the crystal structure of O3[Nb(TPP)]2 was quite different from that for Mo derivatives as shown in Fig. 13 [184]. Oxo-metalloporphyrins other than those described above are the O@Ti complexes. For instance, Guilard et al. reported the transformtion of Ti(Por)Cl to peroxo-Ti(Por) via O@Ti(Por) [189]. The structures of O@Ti and (O2)Ti(Por) were con®rmed by their crystal structures. Oxo-vanadium(IV) porphyrin complex has a stable V@O bond. The oxo ligand of oxo-vanadium(IV) porphyrin is displaced with halides to form bis halide vanadium(IV) porphyrin complex only when it reacts with thionyl halides or oxalyl halide [190]. The bis halide complex reacts with water to reproduce oxo-vanadium(IV) porphyrin complex. The reaction is utilized to synthesize 18O-labeled oxo-vanadium(IV) porphyrin complex [191]. One electron oxidation of oxo-vanadium(IV) porphyrin gives stable oxo-vanadium(IV) porphyrin p-cation radical complex [192], which allows to facilitate systematic studies of the effect of solvent and porphyrin oxidation on m(V@O) vibration as models for compounds I and II. Raman studies shows that the frequency for m(V@O) decreased with increase in solvent acceptor number, consistent with solvent-induced polarization of the m(V@O) bond [193]. The downshift of m(V@O) band upon ligation increased with increasing donor strength of the ligand; OH > imidazolate > imidazole > pyridine > alcohol

Fig. 13. Crystal structures of cis-dioxo-MoVI(TPP), cis-O@MoVI(TPP)(O2) and O3[Nb(TPP)]2

84

Y. Watanabe á H. Fujii

and water. An earlier work reported that the V@O band shifted to higher frequency for 2A1u cations, such as O@V(OEP‡ ), but to lower frequency for 2 A2u cations, such as O@V(TMP‡ ) [194]. However, a recent study demonstrated that the shift of the m(V@O) band is insensitive to radical type, as pointed out for the oxo-ferryl complexes [191]. Oxo-vanadium(IV) porphyrin has unpaired electron in vanadium dxy orbital, which gives sharp EPR signals [195]. ESEEM spectra of oxo-vanadium porphyrins are also reported [196]. Oxo-vanadium(IV) porphyrin p-cation radical is biradical which exhibits a triplet EPR signal at low temperature [192]. Recent study of O@V(OEP‡ )(H2O) showed that the radical complex forms a dimer both in solid and in solution at low temperature and exhibits intramolecular ferromagnetic coupling and intermolecular antiferromagnetic coupling [197].

7 References 1. Dawson JH, Sono M (1987) Chem Rev 87: 1255 2. Watanabe Y, Groves JT (1992) Molecular mechanism of oxygen activation by cytochrome P-450. In: Boyer PD, Sigman DS (eds) The Enzymes, Academic Press, New York, vol 20, p 405 3. Ortiz de Montellano PR (1986) Cytochrome P450, 2nd edn. Plenum Press, New York 4. Dunford HB (1999) Heme Peroxidases, Wiley-VCH, New York, Chap 4 5. Everse J, Everse KE, Grisham MB (1991) Peroxidases in Chemistry and Biology, CRC Press, Boca Raton, vol I, II 6. Shapiro S, Caspi E (1998) Tetrahedron 54: 5005 7. Guengerich FP, Macdonald TL (1993) Sequential Electron Transfer Reactions Catalyzed by Cytochrome P-450 Enzymes. In: Advances in Electron Transfer Chemistry, vol 3, JAI Press, p 191 8. Mansuy D, Battioni P (1993) Dioxygen Activation at Heme Centers in Enzymes and Synthetic Analogs. In: Reedijk J (ed) Bioinorganic Catalysis, Marcel Dekker, New York, Chap 12, p 395 9. Gunsalus IC, Wagner GC (1978) Bacterial P-450cam Methylene Monooxygenase Components: Cytochrome m, Putidaredoxin, and Putidaredoxin Reductase. In: Fleischer S, Packer L (eds) Methods in Enzymology, Academic Press, New York, vol 52, p166 10. Tabushi I, Koga N (1979) Tetrahedron Lett 3681 11. Tabushi I, Koga N (1978) Tetrahedron Lett 5017 12. Tabushi I, Koga N (1979) J Am Chem Soc 101: 6456 13. Felton RH, Owen GS, Dolphin D, Faier J (1971) J Am Chem Soc 93: 6332 14. Groves JT, Nemo TE, Myers RS (1979) J Am Chem Soc 101: 1032 15. Chang CK, Kuo M-S (1979) J Am Chem Soc 101: 3413 16. Matsuura T (1977) Tetrahedron 33: 2869 17. Gunter MJ, Turner P (1991) Coord Chem Rev 108: 115 18. Meunier B (1992) Chem Rev 92: 1411 19. Sheldon RA (1994) Metalloporphyrins in Catalytic Oxidations, Marcel Dekker, New York 20. Collman JP, Zhang X, Lee VJ, Uffelman ES, Braaauman JI (1993) Science 261: 1404 21. Watanabe Y (1997) Model Studies on Heme Monooxygenases. In: Funabiki T (ed) Oxygenases and Model Systems, Kluwer Academic Publishers, Netherlands, vol 19, p 223 22. Dunford HB (1982) Adv Inorg Biochem 4: 41

Characterization of High-Valent Oxo-Metalloporphyrins

23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65.

85

Blumberg WE, Peisach J, Wittenberg BA, Wittenberg AB (1968) J Biol Chem 243: 1854 Wolberg A, Manassen J (1970) J Am Chem Soc 92: 2982 Moss TH, Ehrenberg A, Bearden AJ (1969) Biochemistry 8: 4159 Fajer J, Borg DC, Forman A, Felton RH, Vegh L, Dolphin D (1973) Ann NY Acad Sci 206: 349 Loew G, Herman ZS (1980) J Am Chem Soc 102: 6173 Dolphin D, Forman A, Borg DC, Fajer J, Felton RH (1971) Proc Natl Acad Sci USA 68: 614 Sandusky PO, Salehi A, Chang CK, Babcock GT (1989) J Am Chem Soc 111: 6437 Palaniappan V, Terner J (1989) J Biol Chem 264: 16046 Fujii H (1993) J Am Chem Soc 115: 4641 Schulz CE, Devaney PW, Winkler H, Debrunner PG, Doan N, Chiang R, Rutter R, Hager LP (1979) FEBS Lett 103: 102 Schulz CE, Rutter R, Sage JT, Debrunner PG, Hager LP (1984) Biochemistry 23: 4743 Morishima I, Ogawa S (1978) J Am Chem Soc 100: 7125 La Mar G, de Ropp JS (1980) J Am Chem Soc 102: 395 Thanabal V, La Mar G, de Ropp JS (1988) Biochemistry 27: 5400 Roberts JE, Hoffman BM, Rutter R, Hager LP (1981) J Biol Chem 256: 2118 Roberts JE, Hoffman BM (1981) J Am Chem Soc 103: 7654 Penner-Hahn JE, Ebel KS, McMurry TJ, Renner M, Balch AL, Groves JT, Dawson JH, Hodgson KO (1986) J Am Chem Soc 108: 7819 Hashimoto S, Tatsuno Y, Kitagawa T (1986) Proc Natl Acad Sci USA 82: 2417 Ogura T, Kitagawa T (1987) J Am Chem Soc 109: 2177 Paeng K-J, Kincaid JR (1988) J Am Chem Soc 110: 7913 Chuang W-J, Van Wart HE (1992) J Biol Chem 267: 13293 Yonetani T, Chance B, Kajiwara S (1966) J Biol Chem 241: 2981 Yonetani T (1976) The Enzymes, Academic Press, New York, vol 13 Yonetani T, Schleyer H, Ehrenberg A (1966) J Biol Chem 241: 3240 Lang G, Spartalian K, Yonetani T (1976) Biochim Biophys Acta 451: 250 Sivaraja M, Goodin, DB, Smith M, Hoffman BM (1989) Science 245: 738 Hashimoto S, Teraoka J, Inubushi T, Yonetani T, Kitagawa T (1986) J Biol Chem 261: 11110 Reczek CM, Sitter AJ, Terner J (1989) J Mol Struct 214: 27 Palcic MM, Rutter R, Araiso T, Hager LP, Dunford HB (1980) Biochem Biophys Res Comm 94: 1123 Rutter R, Hager LP, Dhonau H, Hendrich M, Valentine M, Debrunner P (1984) Biochemistry 23: 6809 Egawa T, Miki H, Ogura T, Makino R, Ishimura I, Kitagawa T (1992) FEBS Lett 305: 206 Schonbaum GR, Chance B (1976) The Enzymes. Academic Press, New York, vol 13 Jones P, Middlemiss DN (1972) Biochem J 130: 411 Chance B (1949) J Biol Chem 180: 947 Gouet P, Jouve H-M, Williams PA, Anderson I, Andreoletti P, Nussaume L, Hajdu J (1996) Nature Struct Biol 3: 951 Benecky MJ, Frew JE, Scowen N, Jones P, Hoffman BM (1993) Biochemistry 32: 11929 Ivancich A, Jouve HM, Sartor B, Gaillard J (1997) Biochemistry 36: 9356 Chuang W-J, Heldt J, Van Wart HE (1989) J Biol Chem 264: 14209 Malmstrom BG (1990) Chem Rev 90: 1247 Kitagawa T, Ogura T (1997) Progress in Inorganic Chemistry. Karlin KD (Ed), Wiley, vol 45: 431 Morgan JE, Verkhovsky MI, WikstroÈm (1996) Biochemistry 35: 12235 Ogura T, Takahashi S, Hirota S, Sinzawa-Itoh K, Yoshikawa S, Kitagawa T (1993) J Am Chem Soc 115: 8527 Groves JT, Haushalter RC, Nakamura M, Nemo TE, Evans BJ (1981) J Am Chem Soc 103: 2884

86

Y. Watanabe á H. Fujii

66. Mandon D, Weiss R, Jayaraj K, Gold A, Terner J, Bill E, Trautwein AX (1992) Inorg Chem 31: 4404 67. Penner-Hahn JE, Eble KS, McMurry TJ, Renner M, Balch AL, Groves JT, Dawson JH, Hodgson KO (1986) J Am Chem Soc 108: 7819 68. Hashimoto S, Tatsuno Y, Kitagawa T (1987) J Am Chem Soc 109: 8096 69. Hashimoto S, Mizutani Y, Tatsuno Y, Kitagawa T (1991) J Am Chem Soc 114: 6542 70. Kincaid JR, Schneider AJ, Paeng K-J (1989) J Am Chem Soc 111: 735 71. Groves JT, Watanabe Y (1986) J Am Chem Soc 108: 7834 72. Groves JT, Watanabe Y (1988) J Am Chem Soc 110: 8443 73. Groves JT, Watanabe Y (1986) J Am Chem Soc 108: 7836 74. Yamaguchi K, Watanabe Y, Morishima I (1992) J Chem Soc, Chem Commun 1709 75. Yamaguchi K, Watanabe Y, Morishima I (1993) J Am Chem Soc 115: 4058 76. Dolphin D, Trayler TG, Xie LY (1997) Acc Chem Rev 30: 251 77. Sugimoto H, Tung H-C, Sawyer DT (1988) J Am Chem Soc 110: 2465 78. Bill E, Ding X-Q, Bominaar EL, Trautwein AX, Winkler H, Mandon D, Weiss R, Gold A, Jayaraj K, Hat®eld WE, Kirk M (1990) Eur J Biochem 188: 665 79. Fujii H, Yoshimura T, Kamada H (1996) Inorg Chem 35: 2373 80. Fujii H (1994) Chem Lett 1491 81. Fujii H (1993) Inorg Chem 32: 1993 82. Fujii H, Ichikawa K (1992) Inorg Chem 31: 1110 83. Czarnecki K, Proniewicz LM, Fujii H, Kincaid J (1996) J Am Chem Soc 118: 4680 84. Czarnecki K, Nimri S, Gross Z, Proniewicz LM, Kincaid J (1996) J Am Chem Soc 118: 2929 85. Gross Z, Nimri S (1994) Inorg Chem 33: 1731 86. Czarnecki K, Kincaid J, Fujii H (1999) J Am Chem Soc 121: 7953 87. Fujii H, Yoshimura T, Kamada H (1997) Inorg Chem 36: 6142 88. Collman JP, Sorrell TN, Hoffman BM (1975) J Am Chem Soc 97: 914 89. Ogoshi H, Sugimoto H, Yoshida Z (1975) Tetrahedron Lett: 2289 90. Higuchi T, Uzu S, Hirobe M (1990) J Am Chem Soc 112: 7051 91. Aissaoui H, Ghirlanda S, GmuÈr C, Woggon W-D (1996) J Mol Cat 113: 393 92. Volz H, Holzbecher M (1997) Angew Chem Int Ed Engl 36: 1142 93. Wagenknecht H-A, Woggon W-D (1997) Angew Chem Int Ed Engl 36: 390 94. Jacob GS, Orme-Johnson WH (1979) Biochemistry 18: 2967 95. Ozawa S, Watanabe Y, Morishima I (1994) J Am Chem Soc 116: 5832 96. Jayaraj K, Gold A, Austin RN, Mandon D, Weiss R, Tenner J, Bill E, MuÈther M, Trautwein AX (1995) J Am Chem Soc 117: 9079 97. Yamaguchi K, Watanabe Y, Morishima I (1992) J Cjem Soc Chem Comm 1721 98. Murakami T, Yamaguchi K, Watanabe Y, Morishima I (1998) Bull Chem Soc Jpn 71: 1343 99. Groves JT, Watanabe Y (1986) J Am Chem Soc 108: 7836 100. Groves JT, Watanabe Y (1988) J Am Chem Soc 110: 8443 101. Mizutani Y, Watanabe Y, Kitagawa T (1994) J Am Chem Soc 116: 3439 102. Tsurumaki H, Watanabe Y, Morishima I (1993) J Am Chem Soc 115: 11784 103. Balch AL, Chan Y-W, Cheng R-J, La Mar GN, Latos-Grazynski L, Renner MW (1984) J Am Chem Soc 106: 7779 104. Chin DH, Balch AL, La Mar GN (1980) J Am Chem Soc 102: 1446 105. Chin DH, La Mar GN, Balch AL (1980) J Am Chem Soc 102: 5945 106. Groves JT, Gross Z, Stern MK (1994) Inorg Chem 33: 5065 107. Phillippi MA, Goff HM (1982) J Am Chem Soc 104: 6026 108. Caldewood TS, Lee WA, Bruice TC (1985) J Am Chem Soc 107: 8273 109. Lee WA, Calderwood TS, Bruice TC (1985) Proc Natl Acad Sci USA 82: 4301 110. Groves JT, Gilbert JA (1986) Inorg Chem 25: 123 111. Swistak C, Mu XH, Kadish KM (1987) Inorg Chem 26: 4360 112. Low DW, Winkler JR, Gray HB (1996) J Am Chem Soc 118: 117 113. Hamachi I, Tsukiji S, Shinkai S, Oishi S (1999) J Am Chem Soc 121: 5500

Characterization of High-Valent Oxo-Metalloporphyrins

114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147.

148. 149.

87

Walling C (1975) Acc Chem Rev 8: 125 Poulos TL, Kraut J (1980) J Biol Chem 255: 8199 Erman JE, Vitello LB, Miller MA, Kraut J (1992) J Am Chem Soc 114: 6592 Erman JE, Vitello LB, Miller MA, Shaw A, Brown KA, Kraut J (1993) Biochemistry 32: 9798 Imai M, Shimada H, Watanabe Y, Matsushima-Hibiya Y, Makino R, Koga H, Horiuchi T, Ishimura Y (1989) Proc Natl Acad Sci USA 86: 7823 Martinis SA, Atkins WM, Stayton PS, Sligar SG (1989) J Am Chem Soc 111: 9252 Shimada H, Sligar SG, Yeom H, Ishimura Y (1997) Heme Monooxygenase. In: Funabiki T (ed) Oxygenases and Model Systems, Kluwer Academic Publishers, Netherlands, pp 195±221 (a) Crane BR, Arvai AS, Ghosh DK, Wu C, Getzoff ED, Stuehr DJ, Tainer JA (1998) Science 279: 2121 (b) Crane BR, Arvai AS, Gachhui R, Wu C, Ghosh DK, Getzoff ED, Stuehr DJ, Tainer JA (1997) Science 278: 425 Quillin ML, Arduini RM, Olson JS, Phillips GN Jr (1993) J Mol Biol 234: 140 (a) Poulos TL, Freer ST, Alden RA, Edwards SL, Skogland U, Takio K, Eriksson B, Xuong N-H, Yonetani T, Kraut J (1980) J Biol Chem 255: 575 (b) Finzel BC, Poulos TL, Kraut J (1984) J Biol Chem 259: 13027 Murthy MRN, Reid TJ, Sicignano A, Tanaka N, Rossmann MG (1981) J Mol Biol 152: 465 Miller MA, Shaw A, Kraut J (1994) Nature Struct Biol 1: 524 Phillips SE, Schoenborn BP (1981) Nature 292: 81 (a) Ozaki S, Matsui T, Watanabe Y (1996) J Am Chem Soc 118: 9784 (b) Ozaki S-I, Matsui T, Watanabe Y (1997) J Am Chem Soc 119: 6666 (c) Matsui T, Ozaki S, Watanabe Y (1997) J Biol Chem 272: 32731 Matsui T, Ozaki S, Liong E, Phillips GN, Watanabe Y (1999) J Biol Chem 273: 2838 Ozaki S-I, Yang H-J, Matsui T, Goto Y, Watanabe Y (1999) Tetrahedron Asymmetry 10: 183 Goto Y, Matsui T, Ozaki S-I, Watanabe Y, Fukuzumi S (1999) J Am Chem Soc p 9497 Hill CL, Schardt BC (1980) J Am Chem Soc 102: 6374 Groves JT, Kruper WJ, Haushalter RC (1980) J Am Chem Soc 102: 6375 Hill CL, Hollanderr FJ (1982) J Am Chem Soc 104: 7318 Camenzind MJ, Hollander FJ, Hill CL (1982) Inorg Chem 21: 4301 Schardt BC, Hollander FJ, Hill CL (1982) J Am Chem Soc 104: 3964 Smegal JA, Schedt BC, CL H (1983) J Am Chem Soc 105: 3510 Bortoloni O, Meunier B (1983) J Chem Soc Chem Commun 1364 Bortolini O, Ricci M, Meunier B (1986) Nouv J Chim 10: 39 (a) Ayougou K, Bill E, Charnock JM, Garner CD, Mandon D, Trautwein AX, Weiss R, Winkler H (1995) Angew Chem Int Ed Eng 34: 343 (b) Schappacher M, Weiss R (1987) Inorg Chem 26: 1189 Groves JT, Watanabe Y, McMurry TJ (1983) J Am Chem Soc 105: 4489 Groves JT, Watanabe Y (1986) Inorg Chem 25: 4808 Early works were reviewed by Boucher LJ, Boucher LJ (1972) Coord Chem Rev 7: 289 Fabry TL, Simo C, Javaherian K (1986) Biochim Biophys Acta 160: 118 Yonetaini T, Asakura T (1969) J Biol Chem 244: 4580 Nick RJ, Ray GB, Fish KM, Spiro TG, Groves JT (1991) J Am Chem Soc 113: 1838 Nick RJ, Ray GB, Fish KM, Spiro TG, Groves JT (1991) J Am Chem Soc 113: 1838 (a) Modi S, Saxena AK, Behere DV, Mitra S (1990) Biochim Biophys Acta 1038: 164 (b) Savena A, Modi S, Behere DV, Mitra S (1990) Biochim Biophys Acta 1041: 83 (c) Khan KK, Mondal MS, Mitra S (1996) J Chem Soc, Dalton Trans 1059 (d) Khan KK, Mondal MS, Mitra S (1998) J Chem Soc Dalton Trans 533 Makino R, Uno T, Nishimura Y, Iizuka T, Tsuboi M, Ishimura Y (1986) J Biol Chem 261: 8376 Low DW, Abedin S, Yang G, Winkler JR, Gray HB (1998) Inorg Chem 37: 1841

88

Y. Watanabe á H. Fujii

150. (a) Groves JT, Stern MK (1988) J Am Chem Soc 110: 8628 (b) Czernuszewicz RC, Su YO, Stern MK, Macor KA, Kim D, Groves JT, Spiro TG (1988) J Am Chem Soc 110: 4158 151. (a) De Poorter B, Meunier B (1985) J Chem Soc Perkin Trans II 1735 (b) De Poorter B, Ricci M, Meunier B (1985) Tetrahedron Lett 26: 4459 (c) Robert A, Meunier B (1988) New J Chem 12: 885 (d) Meunier B, Guilmet E, De Carvalho M-E, Poilblanc R (1984) J Am Chem Soc 106: 6668 (e) Guilment E, Meunier B (1984) J Mol Cat 23: 115 (f) Bernadou J, Fabiano A-S, Robert A, Meunier B (1994) J Am Chem Soc 116: 9375 152. (a) Razenberg JASJ, Nolte RJM, Drenth W (1986) J Chem Soc Chem Commun 277 (b) van der Made AW, Nolte RJM, Drenth W (1990) Recl Trav Chim Pays-Bas 109: 537 153. (a) Collman JP, Kodadeck T, Brauman JI (1986) J Am Chem Soc 108: 2588 (b) Collman JP, Brauman JI, Meunier B, Hayashi T, Kodadeck T, Raybuck SA (1985) J Am Chem Soc 107: 2000 (c) Collman JP, Brauman JI, Hampton PD, Tanaka H, Bohle DS, Hemnre RT (1990) J Am Chem Soc 112: 7980 154. Groves JT, Lee J, Marla SS (1997) J Am Chem Soc 119: 6269 155. Collins TJ, Gordon-Wylie SW (1989) J Am Chem Soc 111: 4511 156. Collins TJ, Powell RD, Slebodnick C, Uffelman ES (1990) J Am Chem Soc 112: 899 157. Workman JM, Powell RD, Procyk AD, Collins TJ, Bocian DF (1992) Inorg Chem 31: 1548 158. (a) Groves JT, Kruper WJJ (1979) J Am Chem Soc 101: 7613 (b) Groves JT, Haushalter RC (1981) J Chem Soc Chem Commun 165 159. Fujii H, Yoshimura T, Kamada H (1997) Inorg Chem 36: 1122 160. Budge JR, Gatehous BMK, Nesbit MC, West BO (1981) J Chem Soc, Chem Commun 370 161. Groves JT, Kruper WJJ, Haushalter RC, Butler WM (1982) Inorg Chem 21: 1363 162. Buchler JW, Lay KL, Castle L, Ullrich V (1982) Inorg Chem 21: 842 163. Yamaji M, Hama Y, Miyazaki Y, Hoshino M (1992) Inorg Chem 31: 932 164. Greager SE, Murray RW (1985) Inorg Chem 24: 3824 165. (a) Siddall TL, Miyaura N, Huffman JC, Kochi JK (1983) J Chem Soc Chem Commun 1185 (b) Samsel EG, Srinivasan K, Kochi JK (1985) J Am Chem Soc 107: 7606 166. Holm RH (1987) Chem Rev 87: 1401 167. (a) Che C-M, Lau K, Lau T-C, Poon C-K (1990) J Am Chem Soc 112: 5176 (b) Leung W-H, Che C-M (1989) J Am Chem Soc 111: 8812 (c) Che C-M, Wong K-Y, Poon C-K (1985) Inorg Chem 24: 1797 (d) Che C-M, Lai T-F, Wong K-Y (1987) Inorg Chem 26: 2289 (e) Che C-M, Leung W-H (1987) J Chem Soc, Chem Commun 1376 168. (a) Marmion ME, Takeuchi KJ (1988) J AM Chem Soc 110: 1472 (b) Acquaye JH, Muller JG, Takeuchi KJ (1993) Inorg Chem 32: 160 169. (a) Bailey AJ, Grif®th WP, White AJP, Williams DJ (1994) J Chem Soc Chem Commun 1833 (b) El-Hendawy AM, Grif®th WP, Piggott B, Williams DJ (1988) J Chem Soc Dalton Trans: 1983 170. Masuda H, Taga T, Osaki K, Sugimoto H, Mori M, Ogoshi H (1981) J Am Chem Soc 103: 2199 171. Collman JP, Barnes CE, Brothers PJ, Collins TJ, Ozawa T, Gallucci JC, Ibers JA (1984) J Am Chem Soc 106: 5151 172. Leung T, James BR, Dolphin D (1983) Inorg Chim Acta 79: 180 173. Groves JT, Quinn R (1984) Inorg Chem 23: 3844 174. Groves JT, Quinn R (1985) J Am Chem Soc 107: 5790 175. James BR (1992) Catal Org React 47: 245 176. Leung W-H, Che C-M (1989) J Am Chem Soc 111: 8812 177. Che C-W, Chung W-C, Lai T-F (1988) Inorg Chem 27: 2801 178. Dubourdeaux P, TrvareÂs M, Grand A, Ramasseul R, Marchon J-C (1995) Inorg Chem Acta 240: 657 179. Leung W-H, Che C-H, Yeung C-H, Poon C-K (1993) Polyhedron 12: 2331 180. Tokita Y, Yamaguchi K, Watanabe Y, Morishima I (1993) Inorg Chem 32: 329 181. Groves JT, Ahn K-H (1987) Inorg Chem 26: 3833 182. Ohtake H, Higuchi T, Hirobe M (1992) J Am Chem Soc 114: 10660

Characterization of High-Valent Oxo-Metalloporphyrins

89

183 (a) Buchler JW, Rohbock K (1972) Inorg Nucl Chem Lett 8: 1073 (b) Buchler JW, Puppe L, Rohbock K, Schneehage HH (1973) Chem Ber 106: 2710 184. Johnson JF, Scheidt WR (1978) Inorg Chem 17: 1280 185. Colin JBC, Cian AD, Weiss R (1983) Angew Chem Int Ed Engl 22: 247 186. Diebold T, Chevrier B, Weiss R (1979) Inorg Chem 18: 1193 187. Tachibana J, Imamura T, Sasaki J (1993) J Chem Soc Chem Commun 1436 188. Mentzen BF, Bonnet MC, Ledon HJ (1980) Inorg Chem 19: 2061 189. Guilard R, Latour J-M, Lecomte C, Marchon J-C, Protas J, Ripoll D (1978) Inorg Chem 17: 1228 190. Richard P, Poncet JL, Barbe JM, Guilard R (1982) J Chem Soc Dalton Trans 1451 191. Czarnecki K, Proniewicz LM, Fujii H, Ji D, Czernuszewicz RS, Kincaid J (1999) Inorg Chem 38: 1543 192. Lemtur A, Chakravorty BK, Dhar TK, Subramanian J (1984) J Phys Chem 88: 5603 193. Oliver Su Y, Czernuszewicz RS, Miller LA, Spiro TG (1988) J Am Chem Soc 110: 4150 194. Macor KA, Czernuszewicz RS, Spiro TG (1990) Inorg Chem 29: 1996 195. Chasteen ND (1981) Biological Magnetic Resonance. In: Berliner LJ, Reuben J (eds) Plenum New York, vol 3, p53 196. Fukui K, Ohya-Nishiguchi H, Kamada H (1993) J Phys Chem 97: 11858 197. Schulz CE, Song H, Lee YJ, Mondal JU, Mohanrao K, Reed CA, Ann Walker F, Scheidt WR (1994) J Am Chem Soc 116: 7196

Characterization, Orbital Description, and Reactivity Patterns of Transition-Metal Oxo Species in the Gas Phase Detlef SchroÈder1, Helmut Schwarz2, Sason Shaik3 1,2 3

1 2 3

Institut fuÈr Organische Chemie der Technischen UniversitaÈt Berlin, Straûe des 17. Juni 135, D-10623 Berlin, Germany Department of Organic Chemistry and the Lise-Meitner-Minerva Centre for Computational Quantum Chemistry, The Hebrew University of Jerusalem, 91904 Jerusalem, Israel E-mail: [email protected] E-mail: [email protected] E-mail: [email protected]

Metal-oxo and -peroxo species play key roles in catalytic oxidations using transition metals. This contribution addresses gas-phase studies of diatomic metal-oxo species and their properties which turn out to be of prime importance for the understanding of the observed reactivity patterns. A second topic concerns higher transition-metal oxides such as di- and trioxides. For these species there exists a structural dichotomy with the corresponding ƒ M(O2). Most of the metal-based oxidations described dioxygen complexes, e.g., MO2 ƒ! involve crossings between surfaces of different spin as crucial steps, and the violation of spin conservation is proposed to determine the reactivity of the late transition-metal oxides. The ability of transition metals to mediate surface crossings via spin-orbit coupling is introduced as a key aspect in oxidation catalysis which has as yet not been fully appreciated. It is further outlined how these fundamental properties may relate to the properties of metal-oxo catalysts in the condensed phase. Keywords: Gas-phase chemistry, Metal oxides, Oxidation, Spin change

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

92

2

Gas-Phase Studies and Applied Catalysis . . . . . . . . . . . . . . . . .

93

3

Properties and Reactivity of Diatomic Transition-Metal Oxides . . . . . . . . . . . . . . . . . . . .

95

4

Bonding Properties of Metal-Oxo Species . . . . . . . . . . . . . . . .

95

5

Reactivity of Metal-Oxo Species . . . . . . . . . . . . . . . . . . . . . . . . 100

5.1 5.2 5.3 5.4

Thermochemical and Mechanistic Considerations . . . . . . General Considerations for Alkane Hydroxylation . . . . . . Spin Inversion as a Key Aspect in Alkane Hydroxylation by Transition-Metal Oxo Species . . . . . . . . . . . . . . . . . . Miscellaneous Reactions of Transition-Metal Monoxides .

6

High-Valent Transition-Metal Oxides . . . . . . . . . . . . . . . . . . . 109

7

Structural Dichotomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

. . . . . 100 . . . . . 103 . . . . . 104 . . . . . 108

Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

92

D. SchroÈder á H. Schwarz á S. Shaik

8

Transition-Metal Dioxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

8.1 8.2 8.3 8.4 8.5 8.6 8.7

ScO‡ 2 . . . . ‡=o= TiO2 .. VO‡ . . . . . 2 . . . . CrO‡ 2 MnO‡ 2 . . . ‡=o= FeO2 . CoO2-ZnO2

9

Miscellaneous Higher Transition-Metal Oxides . . . . . . . . . . . . 118

10

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

11

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

111 111 113 114 116 116 118

1 Introduction An essential aspect in oxidation catalysis by transition metals [M] is their ability to form metal-oxo, -dioxo, and -peroxo species [1]; here [M] stands for a `bare' or a ligated metal atom. These transition-metal oxides and peroxides exhibit a broad range of reactivity which is determined by: (i) the nature of the metal, (ii) the valence of the metal core, and (iii) the ®rst coordination sphere. In addition, solvent effects, aggregation phenomena, and counter ions in¯uence the activity, substrate speci®city and chemoselectivity of transition-metal oxides in catalysis. In metal peroxides [M]-OOR, the additional substituent R (e.g., R = H, alkyl, silyl, or even another metal center) also plays a decisive role. The manifold of reactivity patterns and properties of metal oxides and peroxides is extremely rich. A tailor-made design of oxidation catalysts requires advanced insight into the bonding schemes, thermochemical parameters, and a detailed knowledge of the kinetics of the elementary steps. These intrinsic properties determine the reactivity of metal oxides and the related peroxides towards oxidizable substrates, thus controlling the reactivity pattern at a molecular level. The associated thermochemical and kinetic parameters de®ne the mechanistic basis for the chemo-, regio-, diastereo-, and even enantioselectivities of metal-based oxidation reactions. In the most simpli®ed approach, only the metal-oxo species themselves are considered, e.g., diatomic metal oxides MO or small oxo- and peroxo species such as metal dioxides MO2, metal-oxygen complexes M(O2), or metal peroxides M(OOR). Gas-phase studies using mass-spectrometric techniques enable to study the reactivity of these metal oxides under well-de®ned conditions [2]. For technical reasons, these small entities are usually charged in these studies, but the reactivity of neutrals can also be probed by mass spectrometry [3]. A particular advantage of this experimental approach is that

Characterization, Orbital Description, and Reactivity Patterns

93

the well-de®ned circumstances allow for direct comparison with accurate theoretical data. In fact, the present stage and level of sophistication of computational chemistry not only provide a reasonable level of credibility, but sometimes also reach the capability of predicting reactivity patterns of small metal-oxo compounds ahead of experimental veri®cation. The present contribution is restricted to mononuclear metal oxides and is divided into two major sections. The ®rst deals with the chemistry of diatomic transition-metal oxides and serves to illustrate the fundamental reactivity patterns of metal-oxo units. The second describes higher metal oxides, such as di- and trioxides; transition-metal peroxides are only addressed in this context if relevant with respect to the reactivity of metal-oxo species. The substrates to be oxidized are mostly limited to alkanes, as their selective activation is rightly viewed as a major challenge in oxidation catalysis. While most of our considerations deal with cationic species, the fundamental aspects may equally well apply for other charge states if the electronic situations and redox properties are properly taken into account. Notwithstanding, our preference for using cationic metal oxides is not unintentional because neutral and, in particular, anionic species are less likely to activate hydrocarbons. This simply re¯ects the well-known fact that hydrocarbons are preferentially activated by electrophiles rather than nucleophiles: the electrophiles could be bare metaloxo cations in the gas phase as well as high-valent metal oxides in acidic solutions or in heterogeneous catalysts with acidic sites.

2 Gas-Phase Studies and Applied Catalysis Before describing our research on fundamental properties of metal-oxo and -peroxo species in the diluted gas phase (typically conducted in the lower 10)6 mbar pressure regime), let us brie¯y outline how these intrinsic properties are connected to the behavior of transition-metal catalysts in applied processes. Considering a metal-oxo species [M]@O as an example, the intrinsic properties of this unit are determined by: (i) the choice of the metal that affects the nature of the metal-oxygen bond and the net charge of the [M]@O unit, (ii) the ligands other than the oxo unit attached to the metal center, and (iii) the formal valence state of the metal. Upon proceeding from microscopic to macroscopic systems, these intrinsic properties are then modi®ed by the local environments (e.g., additional ligands, counterions, coordinated solvent molecules) which determine the reactivity patterns as well as the extended environments (e.g., protein backbones in metallo enzymes or the properties of pores in heterogeneous catalysts) affecting substrate speci®city and chemoselectivity. In particular, the rate-limiting steps may differ between idealized gas-phase studies and those of real catalytic processes. Among the broad variety of conceivable effects, for the sake of brevity let us just address product desorption which is particularly decisive in the oxidation of hydrocarbons. Thus, upon hydroxylation of a hydrocarbon RAH to the corresponding alkanol RAOH, the latter is often preferentially coordinated to a metal center because of its better donor

94

D. SchroÈder á H. Schwarz á S. Shaik

properties; liberation is usually brought about by thermal activation. In contrast, hydroxylation of an alkane in the highly diluted gas phase is often associated with the release of the alkanol formed according to the Eq. (1) because the exothermicity of the oxidation provides a driving force for product desorption: RAH ‡ ‰MŠ@O ! ‰MŠ…RAOH† ! RAOH ‡ ‰MŠ

…1†

In higher pressure regimes, the environment may instead serve as an energy bath, thereby resulting in the formation of long-lived (RAOH)[M] product complexes which can determine the overall rate constants in terms of a product inhibition. In addition, preferential coordination of the alkanol relative to the alkane leads to enhanced residence times for the oxidized product at the active site, thus increasing the risk of overoxidation (see below). These are two of the many factors which give rise to the `pressure gap' between idealized gas-phase studies and real catalysis under typical operating conditions [4]. Furthermore, minute pathways apparent as side reactions in the gas-phase studies as well as the presence of trace impurities may become major factors for the effectiveness and the turnover numbers in applied catalysis. In turn, understanding these side reactions in terms of elementary steps occurring at the molecular level may help to optimize the performance of real catalysts. The properties of the model systems described below may thus allow some extrapolation from the gas-phase studies to more dense matter. It needs to be stressed, however, that intrinsic properties cannot be extrapolated to systems which differ in their fundamental aspects. Thus, with appropriate consideration of the boundary conditions, the properties of a mononuclear metal-oxo species [M]@O can be used to propose modi®cations of real catalysts by ligand effects, additives, co-catalysts, promoters etc. In fortunate cases, even the effects of additional metal atoms close to the active site may be estimated as long as the metal-oxo unit remains intact; e.g., in a mixed binuclear metal oxide of the type [M¢]A[M]@O. The central assumption that the reactive entity itself is essentially unchanged usually holds true in homogeneous catalysis and may also be valid for highly dispersed metal catalysts on more or less unreactive supports. If, however, it comes to the association of metal clusters in which oxo-units are l-bridged to two or more metal centers, the [M]@O unit cannot be regarded as a model system anymore. Instead, the corresponding l-bridged metal-oxide clusters [M]mOn of appropriate valence need to be considered in the gas-phase studies. All extrapolations fail, however, if bulk and cooperative effects predominate. The technically important epoxidation of ethene by molecular oxygen on silver contacts appears as such a case [5], and an appropriate gas-phase mnemonic is hardly conceivable. These considerations shall serve to enable the reader to position properly the role of the gas-phase studies in the context of applied catalysis. Thus, while knowledge of the intrinsic properties of metal-oxo and -peroxo species provides insight into reactivity patterns, activation and passivation mechanisms, etc., a direct translation of gas-phase data to real conditions ± not to

Characterization, Orbital Description, and Reactivity Patterns

95

mention the prediction of better catalysts by engineering the active sites [6] ± is impossible and not intended.

3 Properties and Reactivity of Diatomic Transition-Metal Oxides In 1995, we reviewed the chemistry of cationic transition-metal oxides in the gas phase [2]. Along with updating this review by adding the most salient reports which have appeared lately, we shall concentrate on the fundamental reactivity aspects of transition-metal oxides in the gas phase. In this context, the vast improvement of theoretical methods to describe energetics and reactivity patterns of transition-metal compounds is particularly valuable [7].

4 Bonding Properties of Metal-Oxo Species A major, concept that emerges from the gas-phase studies of transition-metal oxides concerns the role of spin states, and it is the spin multiplicity which acts as a decisive factor in oxidation reactions (see below). Therefore, it is a prerequisite to consider the nature of the metal-oxo bond in some more detail; here, the notation metal oxo refers to a situation in which an oxygen atom, having no other bonding partners, is directly bound to a metal center. Two possible bonding schemes evolve: (i) a low-spin [M]@O species having a formal double bond between the metal and oxygen, and (ii) a diradicaloid high-spin [M] AO situation with a covalent s-bond and resonating p interactions. While this distinction is somewhat simplistic because intermediate bonding schemes are imposed by the d-orbitals [8], the gross classi®cation in terms of low- and high-spin situations provides quite an insight as far as understanding of chemical reactivity is concerned. Bonding patterns are discussed for the metal-oxide cations MO+ mostly of the 3d transition metal series [9±12]. Figure 1 shows the molecular orbital scheme derived for the combination of a 3d-block element with atomic oxygen, and it comprises three different types of valence orbitals [11]: 1. Four r orbitals arise from the combinations of 2s and 2pz of oxygen with 3dz2 and 4s of the metal. Accordingly, 1r is basically the low-lying 2s orbital of oxygen, 2r forms a polarized, covalent r-bond between M and O, 3r is almost a non-bonding orbital with a dominant 3dz2 character, and 4r is the anti-bonding counterpart of 2r which by and large involves the 4s orbital of the metal. 2. The px and py orbitals of oxygen and the 3dxz and 3dyz orbitals of the metal give rise to two sets of perpendicular p orbitals, the bonding combinations 1px and 1py and the anti-bonding counterparts 2px and 2py. 3. The 3dxy and 3dx2 y2 orbitals of the metal do not ®nd symmetry match in the valence orbitals of oxygen and are thus classi®ed as non-bonding d orbitals.

96

D. SchroÈder á H. Schwarz á S. Shaik

Fig. 1. General molecular orbital scheme of a diatomic transition-metal oxide

Let us now begin to occupy the molecular orbitals in this scheme for MO+ cations of the 3d transition-metal series. In the valence space, ScO+ has eight electrons giving rise to a 1S+ ground state with a con®guration 1r2 2r2 1p2x 1p2y arising from the perfect pairing of all electrons. Accordingly, all bonding orbitals are doubly occupied while the non-bonding and anti-bonding orbitals are empty (Fig. 2a). This favorable situation gives rise to a strong bond with a dissociation energy at 0 K of D0(Sc+AO) = 165 kcal/mol [13] (Table 1). For titanium as the next member of the 3d series, one of the non-bonding 1d orbitals is singly occupied resulting in a 2D ground state. As population of the d orbital hardly affects the bond between the metal and oxygen, D0(Ti+AO) = 159 kcal/mol is large. Following this scheme, VO+ which has a 3S+ ground state with a 1r2 2r2 1p2x 1p2y 1d1xy 1d1x2 y2 con®guration and D0(V+AO) = 135 kcal/mol. Considering these con®gurations in terms of the electrons involved in the different bonding blocks (r, p, and d), the bonding in ScO+, TiO+, and VO+ ®nds a simple mnemonic in that of carbon monoxide, for which the molecular orbital scheme is similar to that of ScO+ (Fig. 2a). The ®rst ambiguity occurs for the next member of the series, CrO+, in which the additional electron could either occupy one of the anti-bonding 2p orbitals or the mostly non-bonding 3r, hence giving rise to either 4P or 4S+ states. While CrO+ (4S+) has been characterized spectroscopically [14], theory predicts the 4P or 4S+ states to be very close to each other (ca. 0.1 eV [8, 11, 12]), and a de®nitive state assignment cannot be made for the time being. This uncertainty notwithstanding, the key aspect in the present context is that Hund's rule wins over the Aufbau principle in that the doublet state CrO+ (2D) in which the extra electron is spin-coupled into d-manifold is more than 1 eV less stable than the 4P and 4S+ quartet states. Among other factors [15], the occupation of anti-bonding orbitals reduces the bond strength in CrO+ compared to the early transition metals, i.e., D0(Cr+AO) = 86 kcal/mol. Much as for chromium, for MnO+ too the population of the 2p and 3r manifold is

97

Characterization, Orbital Description, and Reactivity Patterns

Fig. 2a±c. Molecular orbital schemes for the ground states of: a ScO+; b FeO+; c CuO+

Table 1. Experimental bond dissociation energies, D0(M+AO) in kcal/mol, of diatomic

transition-metal oxide cations D0

ScO+ TiO+ VO+ CrO+ MnO+ FeO+ CoO+ NiO+ CuO+ ZnO+ a b c d e f g h i j

164.6a 158.6a 134.9a 85.8a 68.0a 80.0a 74.9a 63.2a 37.4a 38.5a

D0 ‹1.4 ‹1.6 ‹3.5 ‹2.8 ‹3.0 ‹1.4 ‹1.2 ‹1.2 ‹3.5 ‹1.2

D0

YO+ ZrO+ NbO+ MoO+

167.0b 178.9b 164.4b 116.7b

‹4.2 ‹2.5 ‹2.5 ‹0.5

RuO+ RhO+ PdO+ AgO+

87.9c 69.6c 33.7c 28.4c

‹1.2 ‹1.4 ‹2.5 ‹1.2

LaO+ HfO+ TaO+ WO+ ReO+ OsO+ IrO+ PtO+ AuO+

206d 173e 188e 126e 115f 100g 59e 77i ±j

‹4 ‹5 ‹15 ‹10 ‹15 ‹12 ±h ±h

[13] Sievers MR, Chen Y-M, Armentrout PB (1996) J Chem Phys 105: 6322 Chen Y-M, Armentrout PB (1995) J Chem Phys 103: 618 [91] [29] [76], p 50 [77] No error bars given [25] For observations of AuO+, see: Hecq A, Vandy M, Hecq M (1980) J Chem Phys 72: 2876 and Aita CR (1987) J Appl Phys 61: 5182

98

D. SchroÈder á H. Schwarz á S. Shaik

almost degenerate, and the 5P or 5S+ states are very close in energy (0.1 eV [8, 16]). The progressive occupation of anti-bonding orbitals inter alia lowers D0(Mn+AO) to 68 kcal/mol. A simple mnemonic for the bonding situations in CrO+ (4P) and MnO+ (5P) is the NO radical [8]. Iron-oxo species are of prime importance in oxidation catalysis. Moreover, the bonding situation of FeO+ can be regarded as representative for the behavior of the oxo species of the late transition metals [8] and is therefore discussed in more detail. Compared to the ideal situation in ScO+ (1S+), FeO+ cation must account for the presence of ®ve extra electrons. Perfect pairing of the Fe+ (6D) and O (3P) atomic ground states in terms of a [M]@O double bond would give rise to quartet FeO+. High-level ab initio calculations, however, predict a sextet ground state [17], and this state assignment has recently been con®rmed experimentally [18]. Thus, ground state FeO+ (6S+) has a 1r2 2r2 1p2x 1p2y 1d1xy 1d1x2 y2 2p1x 2p1y 3r1 con®guration with ®ve singly occupied orbitals (Fig. 2b). Quartet states arising from any conceivable spin couplings are higher in energy [2, 11, 17]. The most important consequence with respect to chemical reactivity is that FeO+ cation can thus not at all be considered in terms of a double bond, but must rather be described in analogy to the bonding situation in triplet dioxygen. In fact, O2 (3S)) and FeO+ (6S+) molecules share 1r2 2r2 1p2x 1p2y 2p1x 2p1y con®gurations in the bonding blocks, i.e. the 2s orbital(s) of oxygen, three doubly occupied orbitals due to the bonding r- and p-combinations, and two singly occupied anti-bonding p-orbitals. The particular interaction between the atoms is best visualized in terms of the corresponding valence-bond picture [8, 10, 11], i.e., the diatomics O2 (3S)) and FeO+ (6S+) exhibit two resonating 3-electron-2-center p bonds perpendicular to each other (Fig. 3). Thus, ground state FeO+ cation is best described as a high-spin [M] AO diradicaloid species rather than the perfect pairing, lowspin situation [M]@O, which one may anticipate ®rst for a metal-oxo unit. The bonding scheme in FeO+ cation leads to a moderate bond strength, D0(Fe+AO) = 80 kcal/mol, but much more important are the reactivity paradigms which can be derived from the analogy to triplet dioxygen. While

Fig. 3a,b. Valence-bond scheme of the resonating 3-electron-2-center bonds in: a triplet

dioxygen; b FeO+ cation

Characterization, Orbital Description, and Reactivity Patterns

99

the oxidations of almost all organic compounds by O2 are exothermic, organic matter is metastable in oxygen (and air) because kinetic barriers are signi®cant. Moreover, the combustion chemistry occurring after ignition is usually non-speci®c, and selective, partial oxidation is often dif®cult to achieve without catalysis. This behavior can precisely be attributed to the triplet ground state of dioxygen. For example, the initial step in the oxidation of methane by O2 to yield methyl hydroperoxide is exothermic only if a concerted process occurs, e.g., at Eq. (2). In a step-wise process, e.g., at Eqs. (3a,b), however, the hydrogen abstraction in the ®rst step is highly endothermic: CH4 ‡ O2 ! H3 CAOOH;

Dr H ˆ

CH4 ‡ O2 ! CH3 ‡  OOH; CH3 ‡  OOH ! H3 CAOOH;

13:5 kcal/mol

Dr H ˆ ‡55:1 kcal/mol Dr H ˆ

68:6 kcal/mol

…2† …3a† …3b†

The concerted insertion at Eq. (2) is, however, spin-forbidden because methane and methyl hydroperoxide are singlets while O2 has a triplet ground state. Consequently, non-radicaloid oxidations with molecular oxygen experience a spin-inversion bottleneck. In contrast, the stepwise sequence via reactions at Eqs. (3a,b) circumvents the bottleneck, but requires large activation energies which often coincides with the onset of combustion. In fact, many catalysts employed for the partial oxidation of alkanes induce initial homolytic RAH bond cleavages to the corresponding radicals [19]; thus the often observed poor selectivities have their origin in the very ®rst step. As outlined further below, this reactivity scheme applies also to FeO+ cation in that its concerted insertion into an RAH bond to yield the intermediate RAFe+AOH cannot occur unless spin inversion has taken place. Bonding in the remaining MO+ cations of the 3d metals follows the scheme outlined for FeO+ cation. Thus, CoO+ exhibits a 5D state in which one of the d orbitals is doubly occupied. Then, the d-manifold is completely ®lled in NiO+ (4S)), and even in CuO+ the analogy to triplet oxygen persists having a 3S) ground state with 1r22r21p2x 1p2y 1d2xy 1d2x2 y2 2p1x 2p1y 3r2 con®guration (Fig. 2c). Only with ZnO+ does the 2p manifold begin to be ®lled further. The increased occupation of non- and anti-bonding orbitals is associated with decreasing bond strengths, i.e., D0(Fe+AO) = 80 kcal/mol, D0(Co+AO) = 75 kcal/mol, D0(Ni+AO) = 63 kcal/mol, D0(Cu+AO) = 37 kcal/mol, and D0(Zn+AO) = 39 kcal/mol. The slight increase with ZnO+ can be attributed to a more pronounced contribution of ionic con®gurations to the strongly polarized metal-oxygen bond [20]. This bonding scheme of transition-metal oxides is by no means con®ned to the cationic species, and analogous arguments apply for neutral and anionic species as well [9, 21]. For example, the O2 mnemonic of the bonding block holds true not only in FeO+ (6S+), but also in the 5D and 4D ground states of the neutral and the anionic counterparts [22]. Due to the lower oxidation state of the metal, e.g., Fe(II) in neutral FeO and Fe(I) in FeO), the ability of these

100

D. SchroÈder á H. Schwarz á S. Shaik

oxides as oxidizing agents is signi®cantly lowered, however, as might be inferred from the increasing bond strengths, i.e., D0(Fe+AO) = 80 kcal/mol, D0(FeAO) = 101 kcal/mol, and D0(Fe)AO) = 132 kcal/mol [23]. This ordering provides a further rationale for the fact that most of the oxidation reactions described here involve cationic species, because the neutral and anionic counterparts are of lower valence state, and thus weaker oxidizers. For simple non-functionalized hydrocarbons as the substrates of major interest, the lack of reactivity towards nucleophilic reagents further disfavors oxidations with metal-oxide anions. In fact, we are not aware of any CAH bond activation of an alkane by transition-metal oxide anions. Except for comprehensive computational studies of the 4d-metal monoxides [24], much less is known about the properties of 4d- and 5-transitionmetal oxides. Nevertheless, none of the ®ndings reported so far con¯icts with the bonding schemes outlined above. For example, even though the atomic properties of platinum largely differ from its 3d-congener nickel, NiO+ and PtO+ share 4S) ground states with high-spin situations [8, 11, 12, 25]. Note, however, that spin is no longer a good quantum number for 5d elements because relativistic effects are signi®cant [26]. In this context, even the few complete mechanistic schemes which have recently been deduced from highlevel calculations of 5d-metal oxides [25, 27] have to be viewed as ®rst-order approximations because present theoretical methods still do not enable accurate incorporation of spin-orbit coupling in the algorithms used to locate and optimize minima and transition structures in systems of moderate size.

5 Reactivity of Metal-Oxo Species Before addressing the course of alkane oxidation by transition-metal oxo species, some fundamental conditions are to be de®ned which must be met in order to afford an effective catalyst having a metal-oxo unit as active site: (i) the metal-oxo species must be capable of transferring an oxygen atom to a given substrate, and (ii) re-oxidation of the reduced form of the metal needs to be feasible. 5.1 Thermochemical and Mechanistic Considerations

Let us specify these requirements for the metal-mediated hydroxylation of an alkane RAH according to the catalytic sequence given at Eqs. (1) and (4); in the latter, áOñ stands for any putative O-atom donor, e.g., peroxides, ozone, dioxygen etc.: RAH ‡ ‰MŠ@O ! RAOH ‡ ‰MŠ

…1†

‰MŠ ‡ hOi ! ‰MŠ@O

…4†

Characterization, Orbital Description, and Reactivity Patterns

101

The CAH bond dissociation energies of alkanes range from 103 kcal/mol for methane to about 90 kcal/mol for tertiary positions, while the related CAO bonds of the alkanols formed in the reaction at Eq. (1) amount to 90±95 kcal/ mol. For example, D0(H3CAH) = 103.3 kcal/mol [28] vs D0(H3CAOH) = 90.3 kcal/mol [28] in the conversion methane ® methanol compared to D0((H3C)3CAH) = 95.0 kcal/mol [28] vs D0((H3C)3CAOH) = 94.6 kcal/mol [29] for the oxidation iso-butane ® tert-butanol. Accordingly, O-atom transfer from a metal oxide to an alkane is exothermic, if D0(MAO) < [101.4 kcal/ mol + D0(RAOH) ) D0(RAH)], that is D0(MAO) < 88.4 kcal/mol for the hydroxylation of methane (R = CH3) and D0(MAO) < 101.0 kcal/mol for the tertiary CAH bond in iso-butane (R = tert-C4H9). From a thermochemical point of view, reactivity is therefore expected to increase with decreasing D0(MAO). However, if the MAO bond strengths get too low, re-oxidation of the reduced form to the metal-oxo species according to the reaction at Eq. (4) may become dif®cult. For example, CuO+ is expected to ba a potent reagent for CAH bond activation because D0(Cu+AO) amounts to only 37 kcal/mol and the binding pattern suggests a substantial radical-type character on oxygen [9, 30]. However, precisely because of the low af®nity for oxygen, generation of CuO+ from Cu+ in the gas phase under conditions which would permit further reactivity studies with CuO+ has not been achieved so far [31]. From a thermochemical point of view, the bond strengths of those metal-oxo species that are potentially attractive for catalysis thus fall in the range of 70±110 kcal/mol, thereby allowing for hydroxylation of hydrocarbons as well as for facile re-oxidation. Thus, oxidation catalysis meets inter alia two conceptual concerns: (i) the discovery of highly reactive metal-oxo species, and (ii) the development of methods for the re-oxidation of the reduced counterparts. Various concepts are possible to ful®ll these requirements, e.g., separation of substrate oxidation and re-oxidation of the catalyst in time or space. An ideal scenario would, however, involve catalysts which exhibit high speci®cities in either states, i.e., the oxidized forms only react with the substrate and the reduced forms only react with the terminal oxidant. Such procedures would allow for the continuous feed of substrate/oxidant mixtures in a simple manner. Of course, handling these mixtures without hazard poses some technical problems, yet it has been managed in several industrial processes. Another general concern in oxidation catalysis is the risk of overoxidation which results in undesired consumption of material and the production of excessive heat. It is obvious that a metal-oxo species capable of activating alkanes can also afford oxidation of the hydroxylated products. For example, the CAH bonds of methanol are much weaker than those in methane, i.e., D0(HACH2OH) = 94.5 compared to D0(H3CAH) = 103.3 kcal/mol [28]. Moreover, the polar oxidation products are likely to bind more strongly to the active site(s) compared to the hydrocarbon substrate, thereby increasing the residence time of the product on the catalyst and hence favoring overoxidation even further. Two obvious strategies to minimize overoxidation are operation at low conversions and/or addition of moderators. For example, the desorption of a hydroxylation product may be facilitated by adding steam to the feed because H2O may replace ROH ligands at the active site(s).

102

D. SchroÈder á H. Schwarz á S. Shaik

While overoxidation appears as a general problem in oxidation catalysis, let us address one speci®c system in some more detail. Thus, the time-honored Gif-systems developed by the late Sir Derek Barton [32] bring about the oxidation of cyclohexane to cyclohexanone while oxidation of cyclohexanol is less ef®cient compared to that of the cycloalkane precursor [33]. While the precise mechanistic details of the Gif-oxidation are not yet entirely clear, some speculation about a rationalization of this conceptually interesting effect are indicated as they may help to exploit other strategies to prevent overoxidation. Based on the arguments raised above, it is quite obvious that thermochemical parameters cannot account for the preferred alkane oxidation by the Gifsystems, because the CAH bonds of cyclohexane are stronger than the a-CAH bond in cyclohexanol. For the same reason, it appears puzzling that the activation of the weaker CAH bond of the alcohol is kinetically hampered compared with alkane activation, unless there exists a mechanistic switch. There is consensus that the Gif-systems seem to proceed via the attack of the alkane by a metal-oxo species leading to the formation of alkyl radicals as intermediates [34, 35] which are then trapped by molecular oxygen present in the mixture to yield hydroperoxides [32]. Subsequently, the peroxides rearrange to the ketones in terms of the Hock reaction. Whatever the precise mechanistic details of the complex reaction sequence are, let us propose the presence of a second ligand, here a hydroxyl group, at the active site in order to rationalize the surprising preference for oxidation of alkanes vs alkanols. Approach of the hydrocarbon substrate can only yield products if it occurs towards the metal-oxo unit (path a in Scheme 1). The alkanol could, however, either approach the oxo-unit (resulting in oxidation) or react with the additional ligand via substitution (path b in Scheme 1); for a related scenario see [36]. Given the low polarizability of CAH bonds in alkanes, the RAOH dipole may in fact enforce path b for the alkanol. Covalent attachment at the active site may further disfavor CAH bond activation via the metal-oxo unit and thereby can account for the observed selectivity. Similar arguments can be raised for the moderate overoxidation of the keto products in the Gif-systems, e.g., formation of hemiketals with the metal bound hydroxyl ligand. While this scenario is purely speculative, it needs to be stressed that neither thermochemical nor kinetic arguments can account for the disfavored oxidation of alkanols by the Gif-systems. It is the competition with another process in terms of a mechanistic switch (here coordination at a neighboring position)

Scheme 1.

Characterization, Orbital Description, and Reactivity Patterns

103

which may serve to provide a conceptual guidance. So far, the Gif-systems appear to be quite unique in this particular respect, and the mechanistic proposal made here may thus stimulate further research on catalytic systems in which overoxidation experiences a bottleneck. 5.2 General Considerations for Alkane Hydroxylation

Based upon the above consideration of thermochemical criteria, the following generalizations can be drawn for the transition-metal mediated oxidations; note that this comparison is con®ned to metal-oxo units as reactive sites and particularly the reactivity of metal peroxides (e.g., the Sharpless epoxidation) is not covered here. 1. Unless hypervalent, metal-oxo species of early transition metals are unlikely to bring about alkane oxidation because of thermochemical restrictions arising from the large oxophilicities of these elements. In fact, the oxo species of the early transition metals are poorer CAH- and CACbond activators in comparison with the bare metals themselves [37]. This inertness of the early transition-metal oxides ®ts nicely with the CO bonding mnemonic mentioned above. If hypervalent, however, the oxides of these metals are expected to behave as oxygen-centered radicals. These features are re¯ected in the gas-phase properties of the corresponding oxide cations. For example, TiO+ as a formal titanium(III) compound is incapable of oxygenating hydrocarbons; in fact, the opposite can occur, e.g., bare Ti+ reduces water to afford TiO+ + H2 [38, 39]. In marked contrast, the formally hypervalent TiO‡ 2 cation behaves as a typical oxygencentered radical and even abstracts a hydrogen atom from water, D0(HOAH) = 118.1 kcal/mol [28], according to the reaction at Eq. (5) [40]: ‡  TiO‡ 2 ‡ H2 O ! Ti(O)OH ‡ HO

…5†

Radical-type reactions of this kind often have low selectivities and are thus of limited value for the selective partial oxidations; complete oxidation may be useful, however, and the rutile-based photo-oxidative treatment of waste waters falls into this category. 2. Some transition-metal oxides in their highest oxidation states (i.e., d0 compounds) can bring about the oxidation of various organic compounds, e.g., CrO2Cl2, MnO4 , RuO4, and OsO4. However, the organic substrates to be oxidized are either con®ned to reactive ones, e.g., bearing allylic or benzylic CAH bonds, ole®ns etc., or the reactions proceed via radical-type mechanisms [41, 42]. 3. Oxides of the late transition metals form the basis of various powerful oxidation catalysts which are capable of activating alkanes. Moreover, some of these oxidants circumvent mere radical-type pathways and thereby allow for regio- and stereoselective functionalizations of alkanes. The most prominent examples in this respect are cytochrome P-450 and methane

104

D. SchroÈder á H. Schwarz á S. Shaik

monooxygenase [43]; both enzymes share iron cores and afford the oxidation of a broad variety of substrates including methane. Let us therefore focus on the iron based systems which are most relevant and can further be regarded as being representative for the behavior of other oxo species of late transition metals; for a recent survey of the gas-phase chemistry of iron in general, see [44]. 5.3 Spin Inversion as a Key Aspect in Alkane Hydroxylation by Transition-Metal Oxo Species

The conventional view of potential-energy surfaces regards species of different spin multiplicities as being entirely separated from each other; let us term this behavior single-state reactivity (SSR). Here, we will show that the thermal reactivity of metal-oxo species is dominated by interaction of surfaces having different spin. We term this behavior as two-state reactivity (TSR), i.e., the interference of two spin surfaces along the reaction path of a thermally activated process with the crucial feature that the rate-determining step is associated with the spin inversion [45]. For didactic purposes, let us discuss the most simple oxidation brought about by MO+ cations in some more detail. Dating back from 1994, several in-depth gas-phase studies dealt with Eq. (6) in the FeO+/H2 system [46]. Subsequent theoretical studies [9, 10, 47] have demonstrated that this simple reaction may serve to illustrate central features of TSR as well as its role as a key aspect in the reactivity of transition-metal oxides: FeO‡ ‡ H2 ! Fe‡ ‡ H2 O

…6†

Considering the ground state of reactants and products, the reaction at Eq. (6) is formally spin-allowed as Fe+ and FeO+ are sextets and H2 and H2O are singlets. Initiated by some counterintuitive experimental observations, a number of detailed theoretical studies have been performed. To cut a long story short, the lowest-energy path of the reaction at Eq. (6) involves a double spin ¯ip from the sextet to the quartet surface close to the entrance and back to the high-spin surface in the exit channel (Fig. 4). As the potential-energy surface shown in Fig. 4 has been discussed quite comprehensively [9, 10, 47], here we would like to focus on two particular questions: 1. Why is the rate-determining transition structure (TS) of the quartet surface lower in energy than the sextet TS? 2. How is the crossover between different spin multiplicities brought about? The answer to the ®rst question is inherent to the orbital description of the metal-oxo species made above. Thus, the high-spin, sextet ground state of FeO+ has a diradicaloid bonding scheme and cannot react in a concerted manner with the substrate. Bond activation must therefore proceed stepwise and hence involves considerable activation barriers as is the case with

Characterization, Orbital Description, and Reactivity Patterns

105

Fig. 4. Schematic potential-energy surfaces for alkane hydroxylation by FeO+; see [11, 47]

for details. Note that product formation involves two spin inversions to occur along the lowest lying reaction path

oxidations using molecular oxygen (see above). In the excited quartet states, however, bond cleavage and bond formation can occur in concert which in turn lowers the energy demand of the TS. The same arguments apply for the insertion intermediate HAFe+AOH in that iron can form two covalent bonds on the quartet surface, whereas an anti-bonding s-orbital is occupied in the sextet electromer. The fundamentally different reactivity patterns of the high- and low-spin states of metal oxo species can be sketched in terms of a simple spin counting (Scheme 2). Thus, RAH bond insertion of a high-spin state would inevitably lead to a partially anti-bonding interaction, whereas the low-spin state can afford two covalent bonds via perfect pairing. Accordingly, the low-spin insertion intermediates are lower in energy than their high-spin electromers and for precisely the same reasons the insertion barriers are lowered on the low-spin surface. In even a more general sense, the high-spin metal-oxo species can only undergo single-bond reactions, i.e., atom abstractions, radical-type processes, or electron transfer, whereas in addition to these, the low-spin states can also promote two-bond reactions, such as concerted bond insertions, anion transfers, etc. [10]. Note that this classi®cation of the

Scheme 2.

106

D. SchroÈder á H. Schwarz á S. Shaik

reactivity of the metal-oxo species agrees well with the O2 bonding analogy. Thus, the triplet ground state of molecular oxygen is kinetically stable and mostly involved in radical-chain reactions, while concerted processes having low barriers are common for the excited singlet state of dioxygen. TSR is not at all restricted to the particular activation of dihydrogen in the reaction at Eq. (6). Thus, the hydroxylations of methane by FeO+, CoO+, and NiO+ cations [2] involve TSR [47b, 48], the FeO+-mediated oxidation of benzene to phenol [49], and even the hydroxylation of a complex substrate such as norbornane [50] by FeO+ are proposed to occur via TSR [51, 52]. Moreover, Yoshizawa and coworkers have demonstrated that in the hydroxylations of methane by FeOn+ species (n = 0±2) the concerted, low-spin pathways are preferred irrespective of the actual charge of the systems [53]. Accordingly, TSR is not an exception but appears as a general mechanistic pattern for hydroxylations involving oxo species of late transition metals. While spin inversion is a crucial parameter in TSR, these reactions are formally spin-forbidden. In the absence of external ®elds, crossing between surfaces of different multiplicities can be mediated by spin-orbit coupling of the surfaces involved. Even though relativistic effects do not affect the energetics of 3d metals very much, communication between surfaces of different multiplicities via spin-orbit coupling is a signi®cant factor in the reactivity of these metals. Detailed theoretical studies of the reaction at Eq. (6) have demonstrated that the experimentally observed reaction ef®ciency coincides with the probability of spin-inversion between the sextet and quartet surfaces [47a]. Occurrence of TSR has even been proposed in the reactions of early transition metals [38, 54] for which spin-orbit coupling is lowest in the 3d series. Mediation of TSR by spin-orbit coupling has some important implications for the reaction kinetics if surface crossing is rate limiting. In particular, spin inversion via spin-orbit coupling must not obey Arrhenius-type kinetics, because the `residence time' in the crossing region and hence the probability to invert spin is inversely proportional to temperature. Thus, the rate of TSR may decrease at higher temperatures [55], while at the same time the spin-allowed high-spin channels, namely single-state reactivity, can compete effectively [47b]. FeO‡ ‡ CH4 ! Fe‡ ‡ CH3 OH

…7a†

FeO‡ ‡ CH4 ! FeOH‡ ‡ CH3

…7b†

The experimental results obtained for the FeO+/CH4 system [46d] may serve as an example for the unusual energy dependencies of both reaction rates and branching ratios. Thus, the rates of the reactions at Eqs. (7a,b) exhibit rather different energy dependencies (Fig. 5). As outlined in Scheme 2, formation of the closed-shell hydroxylation product methanol, concomitant with generation of Fe+ in the reaction at Eq. (7a), is a two-bond process and thus requires TSR. As the crossing probability decreases with shortening the lifetime of the reactant encounter complex, the Fe+ product channel drops rapidly with increasing collision energy. The reaction at Eq. (7b), however, leads to the

Characterization, Orbital Description, and Reactivity Patterns

107

Fig. 5. Rate constants of the reactions at Eqs. (7a,b) in the FeO+/CH4 couple as a function of

energy; adopted from [46d] to which the reader is referred for further details

release of a CH3 radical and can thus occur via TSR as well as SSR. The latter obeys a regular energy dependence such that the rate constant of this process remains almost constant over quite a large energy range. The competition between both channels inter alia changes in the Fe+/FeOH+ branching ratios from about 30:70 at lowest energies to a minimum of about 2:98 in the range 0.5±1.0 eV and then reaches 50:50 at higher energies. Similarly, pronounced variations in rate constants due to SSR/TSR competition occur in the FeO+/H2 [46] and V+/CS2 [55] systems. Thus, while TSR in itself provides an intriguing mechanistic scenario [45], an even more intriguing deviation from classical behavior evolves in cases in which TSR competes with SSR. Based upon the insight gained in the gas-phase studies of these extremely simpli®ed oxidation `catalysts', we have proposed the competition between single- and two-state reactivity to act as a mechanistic distributor in alkane hydroxylations by cytochrome P-450 [56]. In 1998, Newcomb and coworkers [57] argued that the interplay of SSR and TSR may indeed resolve some of the controversial explanations put forward in order to explain the severe experimental anomalies observed in P-450-mediated oxdiations. Very recently, further evidence for the crucial role of a spin-state crossing effect has been reported by Jin and Groves [58] for oxomanganese(V) prophyrins. Ti‡ ‡ H2 O ! TiO‡ ‡ H2

…8†

Interestingly, surface crossing is also observed in the reactions of early transition-metals, but here it occurs in an opposite sense. For example, the

108

D. SchroÈder á H. Schwarz á S. Shaik

exothermic reaction at Eq. (8) of ground state Ti+ (4F) with water yields TiO+ (2D) concomitant with neutral dihydrogen, and thus a net change results from crossing the quartet to the doublet surface [38, 59]. Although the high-spin transition structure has not been located [39], it is safely assumed to be more energy demanding than the low-spin TS. Thus, the reaction at Eq. (8) ®ts the TSR scheme nicely, i.e., a surface crossing en route from reactant complex to the lowest-lying transition structure. Substantial variations of the energydependent cross sections [38] suggest that in the reaction at Eq. (8) spin inversion is also a rate limiting factor. Due to the high oxophilicity of lowvalent titanium, however, oxygen transfer occurs from the substrate to the metal, i.e., a reduction of the substrate rather than an oxidation takes place. 5.4 Miscellaneous Reactions of Transition-Metal Monoxides

Before closing this section, let us brie¯y address other achievements made since the publication of the 1995 review [2]. In the 3d series, reactivity studies mostly focused on FeO+ by extending the range of substrates to alkanols [60], ole®ns [61], substituted benzenes [62], heterocycles [63], as well as main group hydrides [64]. While we refer the reader to the original sources for further details, some aspects are noteworthy in the present context. As expected from the cursory discussion above, the oxidation of alkanols by FeO+ proceeds more rapidly than that of the parent hydrocarbons. Interestingly, the reaction mechanisms differ fundamentally from those observed with the alkanes in that initial coordination of the metal-oxide cation to the hydroxy group determines the fate of the reactions. For example, a major pathway in the reactions of some alkanols with FeO+ constitutes a direct hydroxide ion abstraction to afford the corresponding carbocations and neutral Fe(O)(OH); nevertheless, CAH bond activation can still take place to some extent [60]. More extreme, phenol avoids CAH bond activation as an initial step [62b] even though FeO+ is capable of hydroxylating benzene [49]. Mass spectrometric studies in conjunction with extensive 2H and 13C labeling reveal that OAH bond activation of phenol predominates to afford C6H5OAFe+AOH as central intermediate from which the subsequent products evolve. Initial occurrence of ring hydroxylation by FeO+ as a major path is rigorously excluded by 18Olabeling [62b]. Finally, the reactions of FeO+ with aromatic amines are initiated by electron transfer (ET) from the amine to the metal oxide [62a]; this result is in close analogy to the ET driven oxidative dealkylation of Nalkylanilines by cytochrome P-450. A few other reactivity studies have been performed with 3d-metal monoxides and organic substrates since 1995 [65]. Signi®cant progress has been achieved for a series of small inorganic reagents. Thus, Sievers and Armentrout examined the reduction of CO2 ® CO by various early transition metals in quite some detail [66]; a highlight concerns the study of the V+/CO2 system in which it was possible to map out crucial parts of the potential-energy surface experimentally [67]. Also noteworthy is the reaction of FeO+ with NO2 [46c] as one of the few examples in which the metal is further oxidized,

Characterization, Orbital Description, and Reactivity Patterns

109

yielding NO+ together with neutral FeO2, i.e., a formal iron(IV) compound (see below). In this context, the ongoing studies of Plane and coworkers on the role of iron oxides in atmospheric chemistry are worth mentioning [68]. Several anionic and neutral transition-metal oxides have been examined recently by different means of optical spectroscopy [22, 69]. Andrews and coworkers have undertaken systematic investigations of the metal-oxide chemistry in doped rare gas matrices; for example, neutral and cationic OM(CO)+/o complexes (M = Ti, V [70]) as well as OM(CO)o/) species (M = CrACu [71]) were examined by matrix-isolation spectroscopy. Among the 4d and 5d series, few reactivity studies with oxidizable substrates have been reported. The reactivities of neutral metal oxides towards alkanes have been examined, but in these studies only the depletion of the reactants, namely MO, has been monitored while the product structures remain undetermined [72]. Oxygen-atom transfer to alkanes does not occur with MoO+ [73], HfO+ [74], TaO+ [75], WO+ [74], ReO+ [74, 76], and OsO+ [77] which is in accordance with their large bond-dissociation energies (Table 1). MoO+ can, however, bring about oxidation of methanol to formaldehyde [78]. Interestingly, ReO+ and OsO+ are even capable of activating hydrocarbons including methane [74, 77], but instead of O-atom transfer, dehydrogenation takes place presumably to yield bisligated oxometal carbenes according to the reaction at Eq. (9): MO‡ ‡ CH4 ! M…O†…CH2 †‡ ‡ H2

…M ˆ Re; Os†

…9†

Particularly noteworthy is a rather extensive theoretical study of the reaction paths in the Pt+/CH4/O2 system [25, 79] which allows for a catalytic oxidation of methane by dioxygen [80]. In this context, the recently reported catalytic sequences for gas-phase oxidations with Ptn On cluster anions are of interest [81]; note, however, that the oxidizable substrate chosen was again not an alkane but carbon monoxide, which is more likely to be attacked by anionic species. Finally, an interesting observation in the series of lanthanide monoxides should be mentioned [82]. Although oxygen-atom transfer cannot be achieved with lanthanides due to their large oxophilicities, the LnO+ cations are much more capable of inducing ole®n oligomerizations than their formally isovalent MX‡ 2 homologs (X = F, OH, Cl, OCH3). This difference demonstrates the decisive role of the metal-oxo unit and has been attributed to the particular reactivity of this unit in terms of intermediate electron transfer as a ratedetermining step [82]. More recently, Gibson used an elegant device to extend these studies to several cationic actinide monoxides and obtained similar results [83].

6 High-Valent Transition-Metal Oxides So far, the discussion has, by and large, been con®ned to metal-monoxide cations, i.e., compounds with formal metal(III) oxidation states. Most applied transition-metal based oxidants exhibit higher oxidation states in their active

110

D. SchroÈder á H. Schwarz á S. Shaik

forms; often these are termed high-valent metal-oxo species. Typical examples include chromium(VI) compounds such as CrO2Cl2, CrO3, and CrO24 , permanganate, Mo, and Ru oxides, and the time honored osmium(VIII) tetroxide. Here we review the present status of the gas-phase experiments conducted with high-valent metal oxides. Note that metal peroxides are only considered if directly relevant to the chemistry of metal-oxo species; in particular, neither the Sharpless epoxidation nor the reactions mediated by methyltrioxorhenium are discussed because they do not involve metal-oxo units in the decisive oxidation steps.

7 Structural Dichotomy If more than one oxygen atom is attached to a metal center, a structural dichotomy results. Thus, for the elemental composition [M,O2] three fundamentally different bonding situations exist: (i) an end-on metal dioxygen complex I, (ii) a side-on complex II, and (iii) an inserted metal dioxide III (Scheme 3). Depending on the spin coupling, these structures possess different bonding mnemonics. For example, high-spin coupled II can be regarded as a mere complex of the metal with triplet dioxygen, whereas its low-spin congener is best described as a covalently bound metal peroxide. Similar arguments apply to [M,On] species with n > 2. These different types of bonding have been distinguished in a comparative study of oxide cations of chromium, iron, and rhenium [84]. For [Cr,O2]+ structures I and III coexist as well-separated minima on different spin surfaces, i.e., the end-on dioxygen complex Cr(O2)+ (6A¢¢) and the dioxide 2 + cation CrO‡ 2 ( A1), respectively [85]. In the case of [Fe,O2] , sextet states of structures II and III are very close in energy and separated by a rather low barrier [86]. Ionization of the high-valent rhenium peroxide CH3Re(O2)2O [87] leaves the peroxo units intact; thus CH3Re(O2)2O+ cation may serve as a representative for the low-spin, peroxo variant of structure II. As a rough but by no means general guide, structure III is energetically favored for low oxidation states as well as for those metals which have a preference for the formation of high-valent oxides, while structures I and II prevail for the late transition metals which do not support high oxidation states. This structural dichotomy needs to be kept in mind in the evaluation of the gas-phase properties of [M,On]+/o/) because more than a single isomer and/or state may be generated under the experimental conditions. In fact, coexistence of structural isomers has been demonstrated for anionic [69, 88], neutral [89], cationic [84±86], and dicationic transition-metal oxides [85].

Scheme 3.

Characterization, Orbital Description, and Reactivity Patterns

111

8 Transition-Metal Dioxides Analogous thermochemical criteria as outlined above for alkane oxidation by metal monoxides apply to transition-metal dioxides, except that D0(MAO) is to be replaced by D0(OMAO). If not prevented by a spin-inversion bottleneck, loss of triplet dioxygen needs to be considered as an additional reaction path in evaluating thermochemical stabilities. Thus, a metal dioxide is metastable with respect to the dissociation MO2 ® M + O2, if D0(OMAO)+D0(MAO) < D0(OAO) = 118 kcal/mol. For ionic [M,O2]+/)-species, the additional electrostatic interaction in the M(O2)+/)-complexes provides roughly 10±30 kcal/mol extra stabilization of structures I and II. Unlike the monoxides, the electronic structures of transition-metal dioxides are more dif®cult to categorize because of bending, symmetry breaking, and the structural dichotomy mentioned above. Let us therefore elucidate the bonding patterns as well as selective aspects of the reactivity of transitionmetal dioxides by progressing through the 3d series and only brie¯y addressing the 4d and 5d homologs; for a comprehensive theoretical study of neutral 4d dioxides, see [90]. 8.1 ScO+2

While D0(Sc+AO) = 165 kcal/mol is largest for all 3d monoxide cations, D0(OSc+AO) = 40 kcal/mol [91] is rather low (Table 2). This result is not surprising because scandium cation has only two valence electrons to bind the ®rst oxygen atom and thus no bonding capabilities are left for a second one. Even though no reactivity studies have so far been performed with ScO‡ 2 , the weakness of the OSc+AO bond implies that ScO‡ can act as an ef®cient O2 atom donor. For the very same reason re-oxidation ScO+ + áOñ ® ScO‡ 2 will be problematic, therefore rendering this ion unattractive as far as oxidation catalysis is concerned. Similar considerations apply for YO‡ with 2 + D0(OY+AO) = 41 kcal/mol and LaO‡ with D (OLa AO) = 23 kcal/mol [91]. 0 2 It would be interesting to know whether these dioxide cations are low- or highspin coupled and whether the oxygen atoms are equivalent, e.g., symmetrical + + ScO‡ 2 vs the asymmetric structures OSc   O and ScOO . 8.2 TiO+/o/) 2

Neutral titanium has four valence electrons and can thus precisely saturate the demands of two oxo ligands. Indeed, neutral TiO2 undergoes perfect pairing resulting in a 1A1 singlet ground state with D0(OTiAO) = 144 kcal/mol; the latter value is much too high in the context of alkane oxidation [40, 92]. Due to the electron-withdrawing oxo ligands, TiO2 has a sizable electron af®nity as well as ionization energy, i.e., EA(TiO2) = 1.6 eV [69f ] and IE(TiO2) = 9.5 eV

112

D. SchroÈder á H. Schwarz á S. Shaik

Table 2. Bond dissociation energies, D0(OM+AO) in kcal/mol, of transition-metal dioxide

cations

D0

D0

ScO‡ 2 TiO‡ 2 VO‡ 2 CrO‡ 2

40a 81b 90b 66c

YO‡ 2 ZrO‡ 2 NbO‡ 2 MoO‡ 2

41a 89b 132b 128f

FeO‡ 2

66d

RuO‡ 2 RhO‡ 2

79g 78h

CuO‡ 2

99e

a b c d e f g h i j k l

D0 LaO‡ 2 HfO‡ 2 TaO‡ 2 WO‡ 2 ReO‡ 2 OsO‡ 2 IrO‡ 2 PtO‡ 2 AuO‡ 2

23a 140h 132h 65i 105j 125k 75h ±l

[91] [40] [85] [86] According to ab initio studies, the ion structure is Cu(O2)+, see [104] [66b] Derived from data given in: Norman JH, Staley HG, Bell WE (1968) Adv Chem Ser 72, Gould RF (ed), ACS Washington, p 101 [29] [76] [77] Norman JH, Staley HG, Bell WE (1965) J Chem Phys 42: 1123 For an observation of AuO‡ 2 , see: Aita CR (1987) J Appl Phys 61: 5182

[40]. Ground state TiO2 (2A1) anion [69f, 92, 93] has one extra electron and is best described as a titanium(III) dioxide, i.e., O@Ti AO $ OATi @O. Accordingly, D0(OTi)AO) = 150 kcal/mol exceeds the bond strength in neutral titanium(IV) dioxide; once again this bond energy is much too high to render this oxo species useful in oxidation processes. 2 In marked contrast, TiO‡ 2 ( B2) cation has a moderate bond strength, i.e., + D0(OTi AO) = 81 kcal/mol and behaves as a highly reactive oxidant which not only activates alkanes but even abstracts a hydrogen atom from water ± see the reaction at Eq. (5). This behavior is in accordance with the description of TiO‡ 2 as a hypervalent compound with radical character centered on oxygen [40]; the latter is due to the fact that the three valence electrons of titanium cannot saturate the needs of two oxo ligands. Even though TiO‡ 2 is a potent reagent for bond activation, as an oxygen-centered radical, unselective hydrogen-atom abstraction is facile. This dioxide ion is thus not considered as a valuable reagent for selective oxidation processes. Likewise, the 4d congener ZrO‡ 2 reacts in a radical-like manner [40], and the reactivity patterns of hafnium oxide cations suggest a similar behavior of HfO‡ 2 [74]. Further, neutral and anionic zirconium and hafnium dioxides have been characterized by matrix spectroscopy [93]. Note that most dioxide cations of lanthanide and actinide metals are either of low reactivity or show radical-type behavior [82b, 94].

Characterization, Orbital Description, and Reactivity Patterns

113

8.3 VO+2

The vanadium(III) species VO2 and the neutral vanadium(IV) dioxide VO2 (2A1) are unfavorable candidates for metal-oxo based oxidations, i.e., D0(OV)AO) = 150 kcal/mol and D0(OVAO) = 132 kcal/mol [40, 95]. In marked contrast, the vanadyl cation VO‡ 2 is capable of activating hydrocarbons concomitant with a reduction from formal vandium(V) to vanadium (III). In VO‡ 2 , the four valence electrons of the metal precisely ful®ll the needs of the oxo units giving rise to a 2A1 ground state, sketched as O@V+@O. This perfect pairing situation notwithstanding, the net positive charge and the electron withdrawing oxo ligands enhance the reactivity of VO‡ 2 and result in a moderate bond strength in a range desired for catalysis. The absolute value of D0(OV+AO) is a matter of debate though. Sievers and Armentrout [67] derived D0(OV+AO) = 71 kcal/mol from the onset of highly endothermic oxygenatom transfer from CO to VO+ cation, whereas recent theoretical and experimental results suggest D0(OV+AO) = 90 kcal/mol [40]. As the endothermic reaction studied by Sievers and Armentrout may only provide a lower bound for D0(OV+AO), due to kinetic hindrance of O-atom transfer from the strongly bound CO molecule, we list the larger value in Table 2. Vanadyl cation is capable of oxidizing a variety of substrates [40] beginning with ethane activation according to the reaction at Eq. (10). We restrict the discussion of the oxidation properties of VO‡ 2 to this particular example: ‡ VO‡ 2 ‡ C2 H6 ! V(OH)2 ‡ C2 H4

…10†

The V(OH)‡ 2 cation formed in the reaction at Eq. (10) is a vanadium(III) compound and exhibits a triplet ground state, whereas the other reagents are singlets. Thus, again the oxidation occurs via spin-inversion. However, because ground state VO‡ 2 is a singlet, the course of the reaction fundamentally differs from the TSR scenario outlined above. According to theoretical studies [40], the lowest-lying TS 1/2 for CAH bond activation evolve from the singlet surface of the reactants via the sequence 1 ® 2 ® 3 (Fig. 6). This is obvious from the arguments raised above in the discussion of TSR because the concerted reaction path evolves from the low-spin surface, i.e., the singlet reactants. Hence, bond activation does not require spin inversion and the subsequent crossing to the triplet surface occurs in the product complex and is thus just a bottleneck in product release. Any eventual restrictions in spin inversion occurring after TS 1/2 as the rate-determining step, i.e., in structures 2 and 3, can thus by and large be expressed in terms of the pre-exponential factor in an Arrhenius formalism, while the actual bond activation is expected to follow classical kinetics. Therefore, we may rather classify the reaction at Eq. (10) as SSR followed by spin inversion rather than TSR in which spin inversion is part of the rate-determining step. In fact, there are not many arguments in favor of a situation which we may term as inverse TSR, i.e., ratelimiting crossing from a low- to a high-spin surface along the reaction coordinate, because in most but unusual cases the low-spin TS is expected to

114

D. SchroÈder á H. Schwarz á S. Shaik

Fig. 6. Geometries of stationary points relevant in the reaction at Eq. (10); adopted from [40]

be lower in energy than its high-spin congener if the reactants already have low-spin ground states. 1 ‡ 1 Like VO‡ 2 ( A1), the 4d congener NbO2 ( A1) also exhibits a perfect pairing + ground state, the enhanced D0(ONb AO) = 132 kcal/mol disfavors its application in catalytic oxidation of alkanes [40]. Similarly, the oxophilicity of tantalum does not allow for O-atom transfer with TaO‡ with 2 D0(OTa+AO) = 140 kcal/mol [29]. Indeed, TaO‡ 2 is formed as the terminal product in the stoichiometric coupling of carbon dioxide and methane to ketene [27, 75]. 8.4 CrO+2

Chromium-dioxide cation has a 2A1 ground state, and the moderate bond strength D0(OCr+AO) = 66 kcal/mol allows for the oxidation of various hydrocarbons [85]. In addition to the ground state dioxide CrO‡ 2 , the existence of the dioxygen complex Cr…O2 †‡ as an isomeric species arising from the ground states of Cr+ (6S) and O2 (3Sg) has been demonstrated in this case; circumstantial evidence for a quartet species has also been obtained. Chromium dioxide is one of the most powerful gas-phase oxidants examined so far. Thus, CrO‡ 2 activates various substrates including dihydro+ gen and methane. Moreover, complete reduction from CrO‡ 2 to Cr is observed to some extent, i.e., transfer of two oxygen atoms to the substrate. An example

Characterization, Orbital Description, and Reactivity Patterns

115

is the reaction at Eq. (11) occurring with methane; the nature of the neutral molecule(s) formed is unknown, but water and formaldehyde are a likely product combination which ®nds support in the inter alia observed formation of Cr(CH2O)+ from the O‡ 2 /CH4 couple: ‡ CrO‡ 2 ‡ CH4 ! Cr ‡ ‰C; H4 ; O2 Š

…11†

Cr‡ ‡ O2 ! Cr…O2 †‡ ! CrO‡ 2

…12†

In this respect, the possible sequence depicted in the reaction at Eq. (12) is also of interest. Though it involves a net change from the sextet to the doublet surface, spin inversion in the reaction at Eq. (12) appears to occur with a ®nite probability in the gas phase, because under multiple collision conditions, Cr…O2 †‡ was found to end up as CrO‡ 2 [85]. Thus, combination of the reactions at Eqs. (11) and (12) is a possible strategy for catalytic oxidations applying dioxygen as a terminal oxidant. While it may appear counterintuitive that a low valent metal, i.e., Cr+ cation, can promote oxidations, this is precisely what has been observed by Bakac and Espenson in the chromium(II)-mediated oxidations of alkanols by dioxygen in protic solutions [96]. Unfortunately, CrO‡ 2 is too reactive in that its reactions with higher alkanes show little selectivities giving rise to a manifold of products. Selectivity poses quite a dilemma in designing reactive oxidants, because those species which are capable of activating the reasonably strong CAH bonds in non-polar substrates such as alkanes are often capable of activating any bond in the substrate as well as in the oxidation products. In fact, high reactivity is often associated with reduced selectivity, and for this very reason intrinsic reactivity needs to be modi®ed by the local environment in order to improve selectivity while keeping reactivity high. For the description of the bonding pattern of CrO‡ 2 , let us refer to the 4d ‡ homolog MoO‡ 2 which has been analyzed in some detail [73a]. In the MoO2 (2A1) ground state, the combination of Mo with two O atoms gives a set of doubly occupied s- and p orbitals which are all bonding except for a doubly occupied orbital of b1 symmetry which, by and large, comprises the 2px orbitals of oxygen and is thus referred to as non-bonding. The uncoupled electron resides in a r orbital of a1 symmetry with slightly anti-bonding character. Thus, the bonding situation of MoO‡ 2 can be described in terms of two metal-oxygen double bonds resulting in a formal Mo(V). The balance of the bonding orbitals increases the stability of MoO‡ 2 and leads to a signi®cantly enhanced bond strength compared with chromium, i.e., D0(OCr+AO) = 66 kcal/mol vs D0(OMo+AO) = 128 kcal/mol. In fact, MoO+ can reduce CO2 to CO concomitant with formation of MoO‡ 2 [66b, 73a]. Accordingly, O-atom transfer is less likely to occur with MoO‡ , 2 and along the series of MoO‡ n cations (n = 1±3), the dioxide shows the lowest reactivity towards hydrocarbon substrates [73, 97]. Similarly, WO‡ with 2 D0(OW+AO) = 132 kcal/mol is not attractive as far as catalytic oxidations are concerned [74].

116

D. SchroÈder á H. Schwarz á S. Shaik

8.5 MnO+2

While MnO2 is a well-known oxidant in the condensed phase, gas-phase oxidations with neutral and cationic MnO2 have not been reported so far. [Mn,O2]) anions have been made by reacting Mn(CO)n anions with dioxygen [98] and laser desorption of bulk manganese oxides [99], respectively. Interestingly, depending on the mode of ion generation different reactivities towards methanol were found, i.e., the ions formed from the Mn(CO)n /O2 couple are unreactive, whereas ef®cient oxidation to formaldehyde occurs with the ions desorbed from solid manganese oxide. These opposing ®ndings indicate the role of the structural dichotomy mentioned above even for anionic species, i.e., formation of Mn…O2 † in one case and MnO2 in the other. Due to the relevance of manganese in oxidation catalysis, further studies of group 7 metal oxides are very much indicated. For example, the 5d congener ReO‡ 2 activates methane according to Eq. (13); note, that the structure of the ionic product is uncertain [74, 76]: ‡ ReO‡ 2 ‡ CH4 ! ‰Re; C; H2 ; O2 Š ‡ H2

…13†

8.6 FeO+/o/) 2

Experimental studies of transient FeO2 date back to Addison and coworkers in 1965 who examined the thermolysis of volatile iron nitrates [100], and various other reports on this species have appeared more recently [68, 69a, 86, 89b, 99, ‡=o= 101]. Interestingly, ‰FeO2 Š appears as a system for which the dichotomy between dioxygen complex and dioxide structure is particularly pronounced and, in addition, several low-lying states exist. In fact, the variations with respect to electronic structures as well as the richness of accessible structural isomers render iron chemistry one of the challenges for contemporary experimental and theoretical methods. For example, the computational predictions for neutral [Fe,O2] suggest the dioxide structure III is more stable, but the ground-state assignments vary from singlet 1A1 to triplet 3B1, quintet 5B2, and even septet 7A1 [68b, 89b, 101]. Based on their theoretical results, Cao et al. [101a] even doubted that the otherwise accurate and reliable photodetachment experiments with FeO2 [69a] involved ground state ions. Thus, we refrain from a more detailed discussion here, and the interested reader should consult the original references. Two aspects which appear to be settled reasonably well concern the [Fe,O2]+ cation which shows an almost degenerate bonding situation between structures II and III [86]: 6 1. The dioxygen complex Fe(O2)+ (6A1) and the dioxide FeO‡ 2 ( A1) are very close in energy, of the same multiplicity and symmetry, and separated by a rather low barrier (Fig. 7). The facile interconversions of structures II and III is of conceptual interest as it provides a spin-allowed route for the activation of dioxygen. In fact, in the 3d series oxidation of organic ligands

Characterization, Orbital Description, and Reactivity Patterns

117

Fig. 7. Schematic potential-energy surface of the [Fe,O2]+ system; adopted from [86] to

which the reader is referred for further details

L is a unique feature of iron in the reactions of M(L)+ species with dioxygen [102]. Due to the energetic proximity, structures II and III, however, [Fe,O2]+ show an ambivalent behavior in that substrates weakly coordinating to iron undergo oxidation, i.e., a reaction typical for a metal dioxide, while for substrates which favorably bind to bare Fe+, simple ligand displacement occurs to yield Fe(L)+ and O2; this is precisely the behavior expected for a dioxygen complex. 2. The pronounced effects of dynamic electron correlation observed for [Fe,O2]+ is remarkable with respect to the theoretical description of transition-metal oxides. CASSCF calculations with reasonably large basis 6 sets predict Fe(O2)+ (6A1) to be 32 kcal/mol more stable than FeO‡ 2 ( A1), while inclusion of perturbation theory at the CASPT2D level with the same 6 basis sets disfavors Fe(O2)+ (6A1) over FeO‡ 2 ( A1) by 5 kcal/mol. Thus, dynamic correlation changes the relative stabilities by as much as 37 kcal/ mol. Quite obviously, the theoretical treatment of [Fe,O2]+ is far from being complete, and further studies of this fundamental, rather challenging, problem are advised. As a note of caution, we add that the pronounced effect of correlation energy may also affect the accuracy of thermochemical predictions made for transition-metal oxides using the nowadays popular hybrid methods [7]. ‡ Unlike RuO‡ 2 , the 5d congener OsO2 has been studied comprehensively [77], ‡ and OsO2 is capable of oxidizing a broad range of substrates. In particular, a potential route for the OsO‡ 2 -catalyzed oxidation of methane to formaldehyde according to Eq. (14) has been proposed [77]:

CH4 ‡ O2 ! CH2 O ‡ H2 O

…14†

118

D. SchroÈder á H. Schwarz á S. Shaik

8.7 CoO2-ZnO2

So far, gas-phase studies of cobalt and nickel dioxides are limited to anions and neutrals and mostly dealt with their spectroscopic features [69e, 99, 103]. In the case of nickel, isomeric Ni(O2)) and NiO2 ions were distinguished by their photoelectron detachment spectra [69e]. Oxidations of organic molecules have to date only been reported for the anionic species with methanol as a substrate and with moderate to low ef®ciencies [99]. Except for some thermochemical information (Table 2), none of the 4d and 5d congeners has been subjected to detailed structural studies nor have their reactivities towards alkanes been reported. Note however, that cationic iridium and platinum complexes react with dioxygen [74, 80], thus suggesting the intermediate [M,O2]+ species. Copper deserves particular attention because nature has chosen this metal as an alternative to iron for the transport and activation of oxygen in living systems [43]. Interestingly, many of these metalloenzymes have mononuclear copper centers, even though one would not expect ef®cient binding of oxygen with a single copper center due to the ®lled 3d shell. Theoretical studies of neutral Cu(O2) indicate that electron transfer plays an important role in this respect [104], i.e., the binding has a signi®cant contribution of a ‰Cu‡  O2  Š con®guration and thus compares to a superoxide. The potential-energy surface of neutral Cu(O2) is extremely ¯at, however, and a de®nitive assignment as either structure I or II cannot be made with the information available [103]. For the cationic species, experimental evidence for structure III has been provided [105], and the co-existence of isomeric Cu(O2)) and CuO2 anions has been demonstrated recently [88]. Considering the biochemical importance of copper in dioxygen activation, further studies of neutral and ionic copper oxides are anticipated. Gas-phase studies of zinc, cadmium, and mercury dioxides have not been reported, but the example of copper may tell us that chemical intuition, which would exclude dioxides with these elements, is not safe against surprises, and one should thus not generally dismiss these elements as oxidation catalysts.

9 Miscellaneous Higher Transition-Metal Oxides Among the cationic trioxides, gas-phase activations of hydrocarbons have only ‡ ‡ been studied for MoO‡ 3 [73, 97], ReO3 [76], and OsO3 [77] in some detail. Of course the structural variety mentioned above for [M,O2] is even more pronounced for [M,O3]; for M = Mo, Re, and Os these three species can be considered as transition-metal trioxides, however. For example, computational 2 studies predict a C3v-symmetrical structure for ground state MoO‡ 3 ( A2) [73]. + As Mo has only ®ve valence electrons, it cannot satisfy the demands of three oxo units and can thus be described as hypervalent with high radical character on the oxygen atoms. Therefore, it is not surprising that typical radical-type ‡ ‡ reactions occur. For example, in analogy to TiO‡ 2 and ZrO2 cations, MoO3

Characterization, Orbital Description, and Reactivity Patterns

119

abstracts a hydrogen atom from water. Accordingly, MoO‡ 3 is rather reactive and can even activate methane. In line with its description as an oxygen centered radical, the reactions are not very speci®c and H-atom abstraction as well as electron transfer (ET) are major routes observed for MoO‡ 3 . Occurrence of ET is obvious considering IE(MoO3) = 11.7 eV, and similarly one could not expect much speci®ty for the 5d congener WO‡ 3 because IE(WO3) = 12.5 eV is even larger [29]. In contrast, atom abstractions and ET are much less ‡ pronounced for ReO‡ 3 and OsO3 , while both ions bring about the activation of various hydrocarbons [76, 77]. Interestingly, OsO‡ 3 does not activate methane, whereas ReO‡ dehydrogenates methane according to Eq. (15): 3 ‡ ReO‡ 3 ‡ CH4 ! ‰Re; C; H2 ; O3 Š ‡ H2

…15a†

‡ ReO‡ 3 ‡ CH4 ! ‰Re; H4 ; O2 Š ‡ CO

…15b†

Recently, reactivities of some [M,O3]) anions (M = Mn, Fe [99], and Nb [106]) have been examined, but these anions do not even bring about the oxidation of methanol to formaldehyde [107]. The only metal tetroxide cation studied so far is ± of course ± OsO‡ 4 ; not quite unexpectedly, it behaves very much as an oxygen-centered radical and atom abstractions as well as ET prevail [77]. [Fe,O4]) made by chemical ionization of Fe(CO)5/O2 mixtures [86], neutral [V,O4] observed in O2-doped matrices [108], and [Re,On]+ ions (n = 4±8) formed in O2-seeded supersonic expansions appear to have peroxide structures [76, 109]. Finally, some substituted metal-oxo species have been examined in the gas phase, ranging from the iron(IV) compound Fe(O)(OH)+ [110] and the ‡ chromium(VI) species CrOF‡ 3 [111] as well as CrO2 Cl [112] to the molecular cations of the rhenium(VII) compounds CH3ReO3 [113] and CH3Re(O2)2O [84]. Except for the reactions of Fe(O)(OH)+ already described in the earlier review [2], no particular reactivity patterns have been observed which are relevant in the context of alkane hydroxylation.

10 Conclusions Gas-phase studies of metal-oxo species provide a wealth of mechanistic insight into fundamental aspects of oxidation catalysis. In particular, it is the combination of experimental studies with theoretical methods that brings about major progress. In this context, the two-state reactivity paradigm for the reaction of metal-oxo species is of paramount importance. Due to their inherent reactivity, the proximity of various electronic states, vast correlation effects, the role of relativistic effects etc., both experimental and theoretical studies have to apply advanced methods in order to tackle the challenging properties of transition-metal oxides. In a more general sense, we consider close cooperation between various branches of research as being essential to make substantial contributions to

120

D. SchroÈder á H. Schwarz á S. Shaik

this area. This demand for interdisciplinarity ranges from theoretical physics (e.g., surface crossing) via chemical physics (e.g., spectroscopy), physical chemistry (e.g., kinetics and thermochemistry), inorganic chemistry (e.g., bonding properties), physical organic chemistry (e.g., reaction mechanisms) to synthetic inorganic and organic synthesis (e.g., volatile metal compounds, isotopically labeled substrates). Last but not least, direct contact with catalyst research in academia and industry is required in order to de®ne the questions to be answered in the hope of improving the understanding of oxidation catalysis at a molecular level. Acknowledgements. Financial support by the Deutsche Forschungsgemeinschaft, the

Volkswagen Stiftung, the Fonds der Chemischen Industrie, and the Gesellschaft der Freunde der Technischen UniversitaÈt Berlin is gratefully acknowledged. Numerous È LS AG are acknowledged for colleagues from the Bayer AG, BASF AG, and DEGUSSA-HU inspiring discussions and directing our attention to industrial oxidation processes. Furthermore, we appreciate the ongoing and fruitful cooperation with Professor P.B. Armentrout, Salt Lake City, and thank Dr. M. Beyer for sending us a copy of [76].

11 References 1. Bernadou J, Meunier B (1998) Chem Commun 2167 2. SchroÈder D, Schwarz H (1995) Angew Chem 107: 2023; Angew Chem Int Ed Engl 34: 1973 3. (a) Schalley CA, Hornung G, SchroÈder D, Schwarz H (1998) Chem Soc Rev 27: 91; (b) Schalley CA, Hornung G, SchroÈder D, Schwarz H (1998) Int J Mass Spectrom Ion Processes 172: 181 4. (a) Ceyer ST (1990) Science 249: 133; (b) Ceyer ST (1994) In: Proceedings of the Robert A. Welch Foundation, 38th Conference on Chemical Research, Chemical Dynamics of Transient Species, Houston, p 157 5. (a) Jorgensen KA (1989) Chem Rev 89: 431; (b) Jorgensen KA, Hoffmann R (1990) J Phys Chem 94: 3046 6. Thomas JM (1997) Chem Eur J 3: 1557 7. (a) Koch W, Hertwig RH (1998) DFT Applications to transition metal problems. In: Schleyer PvR, Allinger NL, Clark T, Gasteiger J, Kollmann PA, Schaefer HF, Schreiner PR (eds) The encyclopedia of computational chemistry. Wiley, Chichester; (b) Truhlar DG, Morokuma K (eds) (1999) Transition state modeling for catalysis. ACS Symp Ser 721, ACS, Washington 8. Shaik S, Danovich D, Fiedler A, SchroÈder D, Schwarz H. (1995) Helv Chim Acta 78: 1393 9. Yamaguchi K, Takahara Y, Fueno T (1984) In: Smith VH, Schaefer HF, Morokuma K (eds) Applied quantum chemistry. Reidel, Boston, p 155 10. Carter EA, Goddard WA (1988) J Phys Chem 92: 2109 11. Fiedler A, SchroÈder D, Shaik S, Schwarz H (1994) J Am Chem Soc 116: 10,734 12. Sodupe M, Branchadell V, Rosi M, Bauschlicher CW (1997) J Phys Chem A 101: 7854 13. If not mentioned otherwise, for the sake of consistency all metal-ligand bond strengths are taken from the following compilation: Freiser BS (ed) (1996) Organometallic ion chemistry. Kluwer, Dordrecht 14. Dyke JM, Gravenor BWJ, Hastings MP, Morris A (1985) J Chem Phys 89: 4613 15. Schilling JB, Beauchamp JL (1988) Organometallics 7: 194 16. Ryan MF, Fiedler A, SchroÈder D, Schwarz H (1995) J Am Chem Soc 117: 2033

Characterization, Orbital Description, and Reactivity Patterns

121

17. Fiedler A, HrusÏaÂk J, Koch W, Schwarz H (1993) Chem Phys Lett 211: 242 18. Husband J, Aguirre F, Ferguson P, Metz R (1999) J Chem Phys 111: 1433 19. (a) Lunsford J (1995) Angew Chem 107: 1059; Angew Chem 34: 970; (b) Weldon MK, Friend CM (1996) Chem Rev 96: 1391 20. Bauschlicher CW, Partridge H (1998) J Chem Phys 109: 8430 21. Piechota J, Suffczynski M (1997) Z Phys Chem 200: 39 22. Drechsler G, Boesl U, BaÈûmann C, Schlag EW (1997) J Chem Phys 107: 2284 23. Harvey JN, Heinemann C, Fiedler A, SchroÈder D, Schwarz H (1996) Chem Eur J 2: 1230 24. (a) Siegbahn PEM (1993) Chem Phys Lett 201: 15; (b) Siegbahn PEM, Blomberg MRA, Svensson M (1993) J Phys Chem 97: 2564 25. Pavlov M, Blomberg MRA, Siegbahn PEM, Wesendrup R, Heinemann C, Schwarz H (1997) J Phys Chem A 101: 1567 26. PyykkoÈ P (1988) Chem Rev 88: 563 27. SaÈndig N, Koch W (1998) Organometallics 17: 2344 28. Berkowitz J, Ellison GB, Gutman D (1994) J Chem Phys 98: 2744 29. Lias SG, Bartmess JE, Liebman JF, Holmes JL, Levin RD, Mallard WG (1988) J Phys Chem Ref Data, 17: Suppl 1 30. Hippe D, Peyerimhoff SD (1992) Mol Phys 76: 293 31. Rodgers MT, Walker B, Armentrout PB (1999) Int J Mass Spectrom 182/183: 99 32. Barton DHR (1998) Tetrahedron 54: 5805 33. Barton DHR, Boivin J, Gastiger M, Morzycki J, Hay-Motherwell RS, Motherwell WB, Obzalik N, Schwarzentruber KM (1986) J Chem Soc Perkin Trans 1: 947 34. (a) Minisci F, Fontana F, Araneo S, Recupero F, Zhao L (1996) Synlett 119; (b) MacFaul PA, Ingold KU, Wayner DDM, Que L (1997) J Am Chem Soc 119: 10,594 35. For a general debate of this aspect, see: Walling C (1998) Acc Chem Res 31: 155; MacFaul PA, Wayner DDM, Ingold KU (1998) Acc Chem Res 31: 159 36. Kim J, Harrison RG, Kim C, Que L (1996) J Am Chem Soc 118: 4373 37. Jackson TC, Carlin TJ, Freiser BS (1986) J Am Chem Soc 108: 1120 38. Chen Y-M, Clemmer DE, Armentrout PB (1994) J Phys Chem 98: 11,490 39. Irigoas A, Fowler JE, Ugalde JM (1998) J Phys Chem 102: 293 40. Harvey JN, Diefenbach M, SchroÈder D, Schwarz H (1999) Int J Mass Spectrom 182/183: 85 41. (a) Ziegler T, Li J (1995) Organometallics 14: 214; (b) Limberg C, KoÈppe R, SchnoÈckel H (1995) Angew Chem 110: 512; Angew Chem Int Ed Engl 37: 496; (c) Deng L, Ziegler T (1996) Organometallics 15: 3011 42. Mayer JM (1998) Acc Chem Res 31: 441 43. Kaim W, Schwederski B (1991) Bioanorganische Chemie, Teubner, Stuttgart 44. Capron L, Mestdagh H, Rolando C (1998) Coord Chem Rev 178±180: 269 45. Plattner DA (1999) Angew Chem 111: 86; Angew Chem Int Ed 38: 82 46. (a) SchroÈder D, Fiedler A, Ryan MF, Schwarz H (1994) J Phys Chem 98: 68; (b) Clemmer DE, Chen Y-M, Khan FA, Armentrout PB (1994) J Phys Chem 98: 6522; (c) Baranov V, Javahery G, Hopkinson AC, Bohme DK (1995) J Am Chem Soc 117: 12801; (d) SchroÈder D, Schwarz H, Clemmer DE, Chen Y, Armentrout PB, Baranov VI, Bohme DK (1997) Int J Mass Spectrom Ion Processes 161: 175 47. (a) Danovich D, Shaik S (1997) J Am Chem Soc 119: 1773; (b) Filatov M, Shaik S (1998) J Phys Chem A 102: 3835 48. (a) Yoshizawa K, Shiota Y, Yamabe T (1997) Chem Eur J 3: 1160; (b) Yoshizawa K, Shiota Y, Yamabe T (1998) J Am Chem Soc 120: 564; (c) Yoshizawa K, Shiota Y, Yamabe T (1999) J Chem Phys 111: 538 49. (a) SchroÈder D, Schwarz H (1992) Helv Chim Acta 75: 1281; (b) Becker H, SchroÈder D, Zummack W, Schwarz H (1994) J Am Chem Soc 116: 1096 50. (a) Schwarz J, Schwarz H (1995) Helv Chim Acta 78: 1013; (b) Schwarz J (1996) PhD Thesis, TU Berlin D83, Shaker, Aachen 51. Yoshizawa K, Shiota Y, Yamabe T (1999) J Am Chem Soc 121: 147 52. Harris N, Shaik S, SchroÈder D, Schwarz H (1999) Helv Chim Acta 82: 1784

122

D. SchroÈder á H. Schwarz á S. Shaik

53. Yoshizawa K, Shiota Y, Yamabe T (1998) Organometallics 17: 2825 54. Irigoras A, Fowler JE, Ugalde JM (1999) J Am Chem Soc 121: 574 and references cited therein 55. (a) SchroÈder D, Heinemann C, Schwarz H, Harvey JN, Dua S, Blanskby SJ, Bowie JH (1998) Chem Eur J 4: 2550; (b) Rue C, Armentrout PB, Kretzschmar I, SchroÈder D, Harvey JN, Schwarz H (1999) J Chem Phys 110: 7858 56. (a) Shaik S, Filatov M, SchroÈder D, Schwarz H (1998) Chem Eur J 4: 193; (b) Filatov M, Harris N, Shaik S (1999) J Chem Soc Perkin Trans 2: 391 57. Toy PH, Newcomb M Hollenberg PF (1998) J Am Chem Soc 120: 7719 58. Jin N, Groves JT (1999) J Am Chem Soc 121: 2923 59. Clemmer DE, Aristov N, Armentrout PB (1993) J Phys Chem 97: 544 60. SchroÈder D, Wesendrup R, Schalley CA, Zummack W, Schwarz H (1996) Helv Chim Acta 79: 123 61. Schwarz J, Wesendrup R, SchroÈder D, Schwarz H (1996) Chem Ber 129: 1463 62. (a) BroÈnstrup M, SchroÈder D, Schwarz H (1999) Can J Chem 77: 774; (b) BroÈnstrup M, SchroÈder D, Schwarz H (2000) Chem Eur J 6: 91 63. Bakhtiar R, Jacobson DB (1996) J Am Soc Mass Spectrom 7: 938 64. (a) BroÈnstrup M, Kretzschmar I, SchroÈder D, Schwarz H (1998) Helv Chim Acta 81: 2348; (b) BroÈnstrup M, SchroÈder D, Schwarz H (1999) Chem Eur J 5: 1176 65. Seemeyer K, SchroÈder D, Kempf M, Lettau O, MuÈller J, Schwarz H (1995) Organometallics 14: 4465 66. (a) Sievers MR, Armentrout PB (1998) Int J Mass Spectrom 179/180: 103; (b) Sievers MR, Armentrout PB (1998) J Phys Chem A 102: 10,574; (c) Sievers MR, Armentrout PB (1998) Int J Mass Spectrom 185/187: 117; (d) Sievers MR, Armentrout PB (1999) Inorg Chem 38: 397 67. Sievers MR, Armentrout PB (1995) J Chem Phys 102: 754 68. (a) Helmer M, Plane JMC (1994) J Chem Soc Farady Trans 90: 31; (b) Helmer M, Plane JMC (1994) J Chem Soc Farady Trans 90: 395; (c) Plane JMC, Rollason RJ (1996) J Chem Soc Farady Trans 92: 4371; (d) Rollason RJ, Plane JMC (1998) J Chem Soc Farady Trans 94: 3067 69. (a) Fan J, Wang L-S (1995) J Chem Phys 102: 8714; (b) Allen MD, Ziurys LM, Browm JM (1996) Chem Phys Lett 257: 130; (c) Wu H, Desai SR, Wang L-S (1996) J Am Chem Soc 118: 5296; (d) Knight LB, Babb R, Ray M, Banisaukas TJ, Russon R, Dailey RS, Davidson ER (1996) J Chem Phys 105: 10,237; (e) Wu H, Wang L-S (1997) J Chem Phys 107: 16; (f) Wu H, Wang L-S (1997) J Chem Phys 107: 8221; (g) Loock H-P, Simard B, Wallin S, Linton C (1998) J Chem Phys 109: 8980; (h) Li X, Wang LS (1998) J Chem Phys 109: 5264 70. Zhou M, Andrews L (1999) J Phys Chem 103: 2066 71. Zhou, M, Liang B, Andrews L (1999) J Phys Chem A 103: 2013 72. (a) Parnis JM, La¯eur RD, Rayner DM (1994) Chem Phys Lett 218: 544; (b) Parnis JM, La¯eur RD, Rayner DM (1995) J Phys Chem 99: 673; (c) Campbell ML (1998) J Chem Soc Faraday Trans 94: 353 73. (a) Kretzschmar I, Fiedler A, Harvey JN, SchroÈder D, Schwarz H (1997) J Phys Chem A 101: 6252; (b) Kretzschmar I, SchroÈder D, Schwarz H (1997) Int J Mass Spectrom Ion Processes 167/168: 103 74. Irikura KK, Beauchamp JL (1991) J Phys Chem 95: 8344 75. Wesendrup R, Schwarz H (1995) Angew Chem 107: 2176; Angew Chem Int Ed Engl 34: 2033 76. Beyer M. (1996) Dissertation, Technische UniversitaÈt MuÈnchen 77. Irikura KK, Beauchamp JL (1989) J Am Chem Soc 111: 75 78. Fialko EF, Kikhtenko AV, Goncharov VB, Zamaraev KI (1997) J Phys Chem 101: 5772 79. Achatz U, Beyer M, Joos S, Fox BS, Niedner-Schatteburg G, Bondybey VE (1999) J Phys Chem (submitted) 80. Wesendrup R, SchroÈder D, Schwarz H (1994) Angew Chem 106: 1232; Angew Chem Int Ed Engl 33: 1174 81. Shi Y, Ervin KM (1998) J Chem Phys 108: 1757

Characterization, Orbital Description, and Reactivity Patterns

123

82. (a) Cornehl HH, Wesendrup J, Harvey JN, Schwarz H (1997) J Chem Soc, Perkin Trans 2: 2283; (b) Cornehl HH, Wesendrup R, Diefenbach M, Schwarz H (1997) Chem Eur J 3: 1083 83. (a) Gibson JK (1997) Organometallics 16: 4214; (b) Gibson JK (1998) J Am Chem Soc 120: 2683; (c) Gibson JK, Haire RG (1998) J Phys Chem A 102: 10746; (d) Gibson JK (1999) Inorg Chem 38: 165 84. SchroÈder D, Fiedler A, Herrmann WA, Schwarz H (1995) Angew Chem 107: 2714; Angew Chem Int Ed Engl 34: 2517 85. Fiedler A, Kretzschmar I, SchroÈder D, Schwarz H (1996) J Am Chem Soc 118: 9941 86. SchroÈder D, Fiedler A, Schwarz J, Schwarz H (1994) Inorg Chem 33: 5094 87. Herrmann WA, Fischer RW, Scherer W, Rauch MU (1993) Angew Chem 105: 1209; Angew Chem Int Ed Engl 32: 1157 88. Wu H, Desai SR, Wang L-S (1995) J Chem Phys 103: 4363 89. (a) Downs AJ, Greene TM, Gordon CM (1995) Inorg Chem 34: 6191; (b) Andrews L, Chertihin GV, Ricca A, Bauschlicher CW (1996) J Am Chem Soc 118: 467; (c) Cao Z, Duran M, Sola (1997) Chem Phys Lett 274: 411 90. Siegbahn PEM (1993) J Phys Chem 97: 9096 91. Clemmer DE, Dalleska NF, Armentrout PB (1992) Chem Phys Lett 190: 259 92. Walsh MB, King RA Schaefer HF (1999) J Chem Phys 110: 5224 93. Chertihin GV, Andrews L (1995) J Phys Chem 99: 6356 94. Heinemann C, Cornehl HH, SchroÈder D, Dolg M, Schwarz H (1996) Inorg Chem 35: 2463 95. Wu H, Wang LS (1998) J Chem Phys 108: 3310 96. Bakac A, Espenson JH (1993) Acc Chem Res 26: 519 97. Cassady CJ, McElvany SW (1992) Organometallics 11: 2367 98. (a) Fokkens H, Gregor IK, Nibbering NMM (1991) Rapid Commun Mass Spectrom 5: 368; (b) van den Berg KJ, Ingemann S, Nibbering NMM, Gregor IK (1994) Rapid Commun Mass Spectrom 8: 895 99. Oliveira MC, Marcalo J, Vieira MC, Almoster Ferreira MA (1999) Int J Mass Spectrom 185±187: 825 100. Addison CC, Johnson BFG, Logan N (1965) J Chem Soc 4490 101. (a) Cao S, Duran M, SolaÁ M (1997) Chem Phys Lett 274: 411; (b) Kellogg CB, Irikura KK (1999) J Phys Chem A 103: 1150 102. (a) SchroÈder D, Schwarz H (1993) Angew Chem 105: 1493; Angew Chem Int Ed Engl 32: 1420; (b) Boissel P, Marty P, Klotz A, de Parseval P, Chaudret B, Serra G (1995) Chem Phys Lett 242: 157 103. Bauschlicher CW, Langhoff SR, Partridge H, Sodupe M (1993) J Phys Chem 97: 856 104. HrusÏaÂk J, Koch W, Schwarz H (1994) J Phys Chem 101: 3898 105. SuÈlzle D, Schwarz H, Moock KH, Terlouw JK (1991) Int J Mass Spectrom Ion Processes 108: 269 106. Jackson P, Fisher KJ, Willett GD (1999) Int J Mass Spectrom (submitted) 107. Keese RG, Chen B, Harms AC, Castleman AW (1993) Int J Mass Spectrom Ion Processes 123: 225 108. (a) Almond MJ, Atkins RW (1994) J Chem Soc Dalton Trans 835; (b) Chertihin GV, Bare WD, Andrews L (1997) J Phys Chem A 101: 5090 109. Beyer M, Berg C, Albert G, Achatz U, Joos S, Niedner-Schatteburg G, Bondybey VE (1997) J Am Chem Soc 119: 1466 110. SchroÈder D, Schwarz H (1991) Angew Chem 103: 987; Angew Chem Int Ed Engl 30: 991 111. Mazurek U, SchroÈder D, Schwarz H (1998) Coll Czech Chem Comm 63: 1498 112. Walba DM, DePuy CH, Grabowski JJ, Bierbaum VM (1984) Organometallics 3: 495 113. SchroÈder D, Herrmann WA, Fischer RW, Schwarz H (1992) Int J Mass Spectrom Ion Processes 122: 99

Quantum Chemical Studies on Metal-Oxo Species Related to the Mechanisms of Methane Monooxygenase and Photosynthetic Oxygen Evolution P.E.M. Siegbahn1, Robert H. Crabtree2 1 2 1 2

Department of Physics, Stockholm University, Box 6730, S-113 85 Stockholm, Sweden Yale Chemistry Department, PO Box 208107, 225 Prospect St., New Haven, CT 06520-8107, USA E-mail: [email protected] E-mail: [email protected]

During the last few years it has become possible to study enzymatic mechanisms by means of high accuracy quantum chemistry methods. In the present review, examples are given of applications of density functional theory to two mechanisms involving redox active centers. In methane monooxygenase an iron dimer complex oxidizes methane to methanol. A hydrogen abstraction mechanism is suggested involving a high-valent Fe(IV,IV) intermediate, termed Q. In Photosystem II an oxyl radical mechanism is suggested for the formation of O2. Different models of the oxygen evolving cluster are discussed. Spin-state considerations are emphasized. Keywords: Density functional theory, Methane monooxygenase, Photosystem II, Spin-state

crossings, Oxyl radicals

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

125

2

The Spin State Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

126

3

Quantum Chemical Studies . . . . . . . . . . . . . . . . . . . . . . . . . . .

127

3.1 3.2

Methane Monooxygenase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Oxygen Evolving Center of Photosystem II . . . . . . . . . . . .

128 133

4

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

141

5

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

142

1 Introduction Oxidation chemistry is one of the most dif®cult problems within the general area of inorganic and bioinorganic reaction mechanisms. Even in model systems, experimental evidence is often hard both to obtain and to interpret ± the dif®culties only increase on moving to enzymes. At least in methane monooxygenase (MMO), X-ray crystal structures of the enzyme are available as a guide but in Photosystem II (PSII) the structure of the key oxygenevolving tetramaganese cluster (OEC) is still uncertain. These factors make quantum chemical studies an indispensable component of any serious attack on these mechanistic problems. Thanks to the large Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

126

P.E.M. Siegbahn á R.H. Crabtree

improvement of quantum chemical methods in recent years, chemically relevant, reliable results can now be obtained even on paramagnetic clusters of the types found in these enzymes. At the simplest level, such studies can suggest which proposed intermediate structures have energies either so high that they can be eliminated as possibilities or so low that they can best be considered as particularly likely intermediates in the enzyme mechanism. In many cases, a good estimate of the kinetic barriers separating plausible intermediates can also be obtained, so that the plausibility of alternate kinetic pathways can also be compared.

2 The Spin State Problem In addition, quantum chemical studies can sometimes be interpreted to suggest general principles or trends that must apply to any mechanism proposed. As an example of a general point of this sort, it has only been very recently that the problem of alternate spin states has received close attention [1]. Experimental studies of the spectroscopic and magnetic properties of a model complex or enzyme intermediate can normally only give information on the ground spin state of a system. It may be, however, that an unobserved excited spin state of the cluster may prove to have suf®ciently higher reactivity than the ground state that the entire reaction pathway may pass via the excited spin state. Quantum chemistry is one of the few ways of probing this issue, which is discussed below in relation to both MMO and PSII. Experiments and calculations on small unsaturated transition metal complexes have shown that spin-crossings are very common in typical redox reactions like oxidative addition. In fact, a spin-crossing is the rule rather than the exception. In a systematic theoretical study of oxidative addition of H2 to second transition row MHx complexes, 19 of the 24 reactions studied had to go through a spin-crossing [2]. Experimentally, these spin-transitions are known often to be quite ef®cient [3, 4]. A prerequisite for a fast transition is normally that at least one of the crossing potential energy surfaces is rather ¯at [5]. The rate of spin state changes may well be much faster in `real' condensed phase complexes than in the gas phase small molecules considered in this work. One case from the condensed phase where spin-state changes were shown to have a decisive effect was for oxidative addition of linear alkanes to MCp(CO) for cobalt, rhodium and iridium [6, 7]. The difference in relative positioning of the singlet and triplet states was shown to lead to a high reactivity for Rh and Ir, while Co is entirely unreactive. A situation which is less common than a single spin-state crossing is one where two crossings occur. Such a situation would be harder to detect experimentally since the spin of the reactants and products could be the same. One of the ®rst times a reaction of this type was demonstrated was for the case of the insertion of a nickel atom into the OAH bond of water [3]. In this case both the reactant and the product are triplets but the barrier on the triplet surface is exceedingly high. In the region of the molecular complex there is therefore an initial transition to a singlet surface which has a low barrier for

Quantum Chemical Studies on Metal-Oxo Species

127

the insertion. The reaction is completed in the region of the products where the complex returns to the triplet state. Similar situations have more recently been demonstrated by Shaik et al. [8], who emphasized the potential role of differential reactivity of different spin states. In the reaction of gas-phase FeO+ with a CH bond, Shaik, Schwartz and coworkers suggest that the ground high spin state, which correlates with the ground high spin state of the products, has a much higher barrier than does the excited low spin state. They argue that the reaction is slowed by the necessity to undergo a double spin state conversion from high to low and back. Extrapolating to the case of CH activation by P-450, they also suggest that the low spin state can undergo oxene-like reactions in which the FeO group inserts into the CAH bond to give RAFeAOAH directly. This would open up several possible pathways to ROH and could in principle account for some of the apparently contradictory experimental data, where the rebound rate exceeds the maximum values required to consider the radical as having a separate existence. Yoshizawa et al. [9] have suggested a similar explanation for the case of MMO. However, there are also entirely different, more probable, explanations for the contradictory experiments as will be discussed further below.

3 Quantum Chemical Studies It is only in the last 4±5 years that quantum chemical methods have advanced to the point that reliable calculations can begin to be carried out on systems as complicated as enzyme active sites. The most important reason for this development is that Density Functional Theory (DFT) has developed into a much more accurate tool than before. In particular, the introduction of terms depending on the gradient of the density to describe the exchange interaction has proven to improve substantially the accuracy [10, 11]. This improvement, together with the improvement obtained by introducing a few semi-empirical parameters and a part of the Hartree-Fock exchange, led to the B3LYP method which has an accuracy that is not far away from that obtained by the most accurate ab initio methods at a small fraction of the cost [12]. The B3LYP method has been used for most of the DFT studies discussed in the present review. This development has meant that much less effort in the quantum chemical studies has to be spent on reaching suf®cient accuracy. Instead, for an enzyme problem most time is often spent on selecting the appropriate model system for study. For testing of different pathways, model ligands (H2O, NH3, OH, etc.) are normally used, but once the general patterns have been established, it is often possible to move to larger ligands much closer to the experimental ones (e.g., imidazole for His, HCOO for Asp or Glu) to check that the energies remain reasonable. The results of typical DFT calculations consist of an optimized structure and the energy corresponding to that structure. The structure can sometimes be compared with physical data on model compounds or enzyme intermediates. The energy is used to decide if the structure in question can be ®tted into a sensible catalytic cycle. Normally a plausible series of intermediates for an exothermic enzyme reaction must

128

P.E.M. Siegbahn á R.H. Crabtree

have energies in a relatively smoothly descending pattern, so that no one step has an excessive endo-or exothermicity. One also has to attempt to show that no other rearrangement product has a much lower energy that would give rise to a thermodynamic sink for the system. In such a case the catalytic reaction would be brought to a halt by the build-up of the `sink' species. Ideally, the barriers to the key conversion steps should be estimated to check feasibility of the pathway. Recent reviews have been written on the study of enzymes by DFT methods and further details of this type of studies can be found there [13, 14]. 3.1 Methane Monooxygenase

Monoxygenases use O2 as primary oxidant but activate it by reduction by two electrons, so the net conversion introduces one oxygen of O2 into the substrate while the other is reduced to H2O. MMO brings about the selective hydroxylation of a variety of unactivated CH bonds, such as in the conversion of methane to methanol [15±17]. The product methanol is much more sensitive to further oxidation, however, so a chemical system normally gives unselective oxidation to CO2. The hydrophobic pocket of MMO expels the product methanol before the iron cluster is activated for a new catalytic cycle. The enzyme active site contains a dimeric non-heme iron cluster not unlike those found in the oxygen binding protein hemerythrin (Hr) and ribonucleotide reductase (RNR), a protein involved in DNA synthesis [18, 19]. Where a CAH bond to be attacked is suf®ciently weak, as in camphor, a cytochrome P-450 dependent monooxygenase enzyme can be involved [20], as in P-540cam from Pseudomonas putida. In this case, the active site contains a heme ± an iron in a porphyrin ring ± and the mechanism is believed to go via an iron oxo formally considered as Fe(V) but better described as having an Fe(IV)@O structure with one oxidizing equivalent present as an electron `hole' in the heme organic p-system. The reaction is believed to proceed by abstraction of an H atom from the substrate by the oxo group to give an iron hydroxo species and a carbon radical. The radical subsequently abstracts the OH group from iron to give the product. Since the overall reaction in the case of MMO is similar, except that the CAH bond of the substrate is unusually strong (104 kcal/mol), it has generally been assumed that a high valent iron oxo was also a key intermediate in MMO, but that its non-heme structure made it more reactive and able to attack methane. In RNR a similar oxo species has been invoked [19] while in Hr [21] the OAO bond of O2 is never broken because this is a transport protein. The MMO from Methylococcus capsulatus (Bath) consists of three proteins, a hydroxylase (protein A), a reductase (protein C), and a regulatory protein, (protein B). The C component uses NADH as primary reductant that provides the electrons necessary to reduce O2, the B component allows C to feed these electrons to A, and the A component interacts with O2 and carries out the substrate oxidation step. Fe2(II,II), Fe2(II,III), and Fe2(III,III) forms of the

Quantum Chemical Studies on Metal-Oxo Species

129

enzyme are accessible, the ®rst of which interacts with O2 to give an Fe2(III,III) peroxo species. Indeed the Fe2(II,II) form of the A protein is able to carry out one turnover of substrate hydroxylation without the need for proteins B or C. Ê in its isolated The active site iron cluster has an Fe  Fe distance of 3.4±3.5 A forms. A crystal structure [22] shows three bridges between the iron atoms: a glutamate, a hydroxo bridge, and an acetate that is believed to be an artefact of the isolation procedure. By analogy with P-450 an Fe2(IV,IV) oxo species might be involved. Consistent with this proposal, the Fe2(III,III) form is weakly active with H2O2 as substrate [23]. Strong evidence for non-heme iron oxo species have been obtained in relevant model systems [24]. Rather unexpected results were obtained for M. trichosporium MMO with chiral ethane (CH3CHDT) as substrate, where 35% inversion of con®guration occurred and a deuterium kinetic isotope effect (KIE) of 4.2 was observed [25]. The quantitative interpretation of the data was complicated by the apparent requirement for an unreasonably short radical lifetime [26, 27]. Rapid freeze-quench techniques allowed a series of intermediates to be identi®ed, including the key species, denoted Q, that appears to be the active oxidant that attacks methane. MoÈssbauer spectroscopy on Q suggests it is an Fe2(IV,IV) species. As well as discussing the mechanistic data, Feig and Lippard list essentially all the mechanistic variants currently proposed [15]. In 1996 Wallar and Lipscomb [16] and in 1997 Valentine and Lippard [28] reviewed the most recent developments in the area. Theoretically, Yoshizawa et al. [29] used the extended HuÈckel method to look at binuclear iron clusters relevant to MMO. They considered CH activation both by bridging peroxide and terminal oxo intermediates and found the latter to be preferred. A C3v distortion of CH4, earlier suggested by Shestakov and Shilov, [30] was assigned a role in helping to activate CH4. An interaction between a ®lled a1 orbital of a C3v-distorted methane and empty d-orbitals on Fe was considered to be particularly important. In addition, 5-coordination of at least one of the Fe atoms was considered best able to provide a suitable site for the reaction. Yoshizawa et al. [9] later ampli®ed these studies, proposing that the activation of methane CAH bond by the high valent Fe2(l-O)2 diamond core of compound Q involves formation of a Q(CH4) complex with both an FeAC bond and an FeAH bond, followed by a concerted H atom abstraction via a four-centered transition state. In the ®rst DFT study of MMO a quite simple model of the iron dimer complex was used [31]. Instead of the actual glutamates and histidines, hydroxyl and water ligands were used. The number of hydroxyl ligands was chosen to give an overall neutral complex with the desired oxidation states. For MMO there are very good experimental indications that the iron dimer complexes involved are all neutral. For example, normal charge counting on the reduced Fe2(II,II) complex with four carboxylates and two imidazoles leads to a neutral complex. The starting structure for a possible compound Q iron dimer structure was furthermore chosen with two six-coordinated irons and two l-oxo bridges. At the end of the B3LYP optimization this led to two ®vecoordinated irons instead. One of the resulting OH  H2O bridges was then

130

P.E.M. Siegbahn á R.H. Crabtree

replaced by a carboxylate bridge in analogy to the protein complex. This leads to a suggested structure 1 for compound Q.

The optimized structure using ferromagnetic spin coupling has at least one striking feature and this is the quite different FeAO distances of the diamond Ê and 1.77 A Ê and two longer ones core obtained, with two short bonds of 1.74 A Ê and 2.05 A Ê . This was explained by the presence of Jahn-Teller (JT) of 2.00 A distortions with the weak JT axis going through the long FeAO bond on one side of the iron atoms and with an empty site on the other side, leading to 5coordination for the irons. Independently an EXAFS measurement suggested a very similar structure [32]. In addition, EXAFS gave an unusually short FeAFe Ê . The FeAFe distance in the optimized distance for compound Q of only 2.46 A Ê geometry is 2.79 A which is quite short, but not as short as the one measured by EXAFS, which suggests that some feature of the bonding may have been missing in this model. Very similar models to study MMO by DFT methods have also been used recently by Basch et al. [33], yielding similar structures. Still using a simple iron dimer model as in 1, several new possible structures for compound Q, not considered in the previous study, were investigated in a more recent study [34]. In the most interesting of these structures there are two, rather than one, bridging carboxylate. When ferromagnetic coupling was Ê to 2.74 A Ê. used this led to a shortening of the FeAFe distance from 2.79 A When d-functions were added on the oxygens the bond distance decreased to Ê . In the ®nal attempt to make the FeAFe bond still shorter, antiferro2.67 A magnetic spin-coupling was used. This is very much harder to converge but after some effort the proper spin-coupling was achieved. With d-functions on oxygen, the optimization converged to a structure with a short FeAFe distance Ê , fortuitously close to the EXAFS result. The effect of using of 2.47 A Ê. antiferromagnetic spin-coupling is thus a shortening by as much as 0.20 A Further analysis of this effect shows that the spins on the irons have not simply been antiferromagnetically coupled but they have also been substantially reduced in size. It turns out that it is this reduction, leading to low-spin coupling of the individual irons, rather than the antiferromagnetic coupling itself, that causes the reduction in the FeAFe distance. In fact, ferromagnetic

Quantum Chemical Studies on Metal-Oxo Species

131

coupling of two low-spin irons give a nearly identical structure and a very similar total energy as the antiferromagnetically coupled singlet. At the next higher level of modeling, the actual functional groups of the glutamates and the histidines were included in the calculation. Using antiferromagnetic coupling this model converged to the structure in Fig. 1. Ê is reasonably close to the experimental EXAFS The FeAFe distance of 2.54 A Ê , considering the model used and the fact that, for example, distance of 2.46 A relativistic effects are not included. It is important to note that the structures of the simple model, 1, and the more advanced one in Fig. 1 have many similarities, which supports the use of the simple models in the previous study. Even though the symmetry is lower for the larger model and that both irons are 6-coordinated and some of the ligands are nitrogen derived, the JT distortions are, for example, quite similar. The spin distribution is also quite similar, but the asymmetry of the complex causes the two l-oxo oxygens to be different with the oxygen trans to the histidines having a somewhat higher spin, +0.57 compared to )0.37, which should make it more reactive. Since the oxygen with the higher spin is the one pointing towards the substrate pocket in MMO, this could be an important effect. Once a model is decided upon for compound Q, the MMO substrate hydroxylation reaction can also be studied. This reaction has been studied extensively experimentally and several possible mechanisms have been suggested [15, 16]. These mechanisms fall essentially into two categories, radical and non-radical mechanisms. In the radical mechanisms the ®rst step is a hydrogen abstraction from the hydrocarbon, while the non-radical mechanisms suggest a concerted insertion pathway. Starting the theoretical investigation of this reaction with the simple model, 1, a methane molecule was moved towards the empty coordination site of one of the irons. A four-

Fig. 1. Suggested model for compound Q of MMO with two bridging carboxylates. The

structure was obtained using antiferromagnetic spin coupling (1A)

132

P.E.M. Siegbahn á R.H. Crabtree

centered starting structure for the transition state was the initial guess, with one FeACH3 bond and one FeOAH bond forming in a concerted manner. As the optimization proceeded the structure changed and eventually converged to a transition state of a pure hydrogen abstraction type. However, a rearrangement is necessary, where one of the bridging l-oxo groups move out to form a terminal oxo ligand which then abtracts a hydrogen atom from methane. The larger model for compound Q in Fig. 1 is much less ¯exibly coordinated than the previous smaller one. If the doubly bridging carboxylates are kept during the reaction, there are not many places on the complex where a methane molecule could be activated. By far the most plausible place for activation is at the l-oxo bridge trans to the histidines. This is close to the hydrophobic pocket in the protein to which methane can diffuse and where it will ®t. Several spin-states (1A, 3A, etc.) were investigated but no low energy activation of methane was found until the ferromagnetic 9A state was tried. On this surface there is a smooth activation of methane and the fully optimized transition state is shown in Fig. 2. The calculated barrier height for activating methane including promotion and zero-point effects is about 10 kcal/mol. The transition state shown in Fig. 2 is clearly one of an almost pure hydrogen abstraction. This is quite different from the results of other B3LYP studies using different types of models [9]. A major feature of these other

Fig. 2. Optimized transition state for CH4 activation in MMO. The structure was obtained

using ferromagnetic spin coupling (9A)

Quantum Chemical Studies on Metal-Oxo Species

133

models is that they have a net positive charge and are coordinatively unsaturated. One of these models is FeO+, while the other one is more realistic with one four-coordinated and one ®ve-coordinated iron atom and a total charge of +3. With these models a four-centered transition state is found. The reason for this is the large electrostatic attraction obtained between methane and the positive metal complex, which leads to deep minima for molecular methane of 19.5 kcal/mol for FeO+ and of 27.3 kcal/mol for the larger model. No such minimum was obtained for the model in Fig. 2. From the above studies it can thus be concluded that the transition state for activating methane in MMO is most likely one of a pure hydrogen abstraction. This is in line with the conventional rebound mechanism [16] but appears inconsistent with the very short lifetimes measured by radical clocks [26, 27]. As described above, an explanation for this contradiction involving an insertion transition state is considered very unlikely since all higher level models point in the same direction of a pure hydrogen abstraction. Instead, the most likely explanation for the contradictory experiments is the following. Very shortly after the pure hydrogen abstraction transition state is passed, there could be a crossing with the potential surface leading to the methanol product, which is extremely attractive and lacks any additional barrier. In fact, if this crossing occurs before the minimum of the methyl radical is reached, no free radical is likely to be found by the radical clock experiments. In line with this scenario, the methanol surface is found to be quite low in the region after the hydrogen abstraction transition state. No exact crossing point has as yet been determined for MMO since this is a quite dif®cult problem in many dimensions. However, in the case of P-450 Shaik and coworkers have recently suggested a similar reaction path and actually succeeded in determining a spincrossing point which does explain the contradictory experimental data [35]. 3.2 The Oxygen Evolving Center of Photosystem II

The mechanism of dioxygen evolution by the Oxygen Evolving Center of Photosystem II (OEC) in green plants, the mechanistic steps involved and, in the absence of a crystal structure, even the structure of the OEC itself are still uncertain [36]. The OEC itself consists of one calcium and four manganese ions per PSII complex, all essential for oxygen evolution, and one chloride cofactor, removal of which does not completely suppress O2 evolution [37]. An Ê [36a], suggests that EXAFS peak assigned to an Mn  Mn separation of 2.7 A the manganese ions are arranged in two pairs of Mn(l-O)2Mn clusters, a very common structure in inorganic model compounds [36f±h]. The two Mn dimers, the Cl and the Ca are believed to be close to and possibly even linked together in the OEC. Water is believed to bind and undergo oxidation at Mn, but the roles of chloride and calcium are much less well de®ned [36d], chloride usually being considered as a probable ligand for Mn, or Ca, or both. In photosynthetic oxygen evolution, each photon absorbtion event leads to an electron transfer which oxidizes the P680 chromophore of PSII to the P680+ cation radical. This radical in turn oxidizes the OEC by one electron via the

134

P.E.M. Siegbahn á R.H. Crabtree

intermediate oxidation of a tyrosine residue, denoted Tyrz. From a saturating ¯ash experiment, which showed maximal O2 evolution on the third, seventh, and eleventh ¯ashes, Kok et al. [38] proposed a generally accepted scheme in which the OEC accumulates oxidizing equivalents by advancing it in oneelectron steps from an S0 oxidation level through S1, S2, and S3 until at the most oxidized S4 oxidation level, O2 is released and the cluster returns to S0. EPR spectroscopy suggests that alternative states of the cluster are possible at the S2 state because two different EPR signals (the multiline signal at g = 2 and the g = 4.1 signal) are seen [39] depending on the temperature and sample preparation. The invariance of the EXAFS Mn K-edge on going from S2 to S3 suggests the Mn atoms are not oxidized in this step [40a]; contrary conclusions have been reported [40b,c]. In a critical experiment, Messinger and coworkers [41] have shown by H2 18O labeling that, of the two O atoms in S2 and S3 that become O2 after S4, one is in a slowly exchanging site and the other in a very rapidly-exchanging site. The slow exchange site has been considered to be a terminal oxo, and the fast exchange site a water or hydroxo group. In any case a symmetrical Mn@O + O@Mn mechanism seems to be unlikely. Ca-depletion abolishes O2-evolution but reconstitution with Sr partially restores activity and modi®es the EPR multiline spectrum of the S2 state of the Mn cluster, suggesting that Ca may be close to the Mn cluster. Strontium EXAFS experiments on PSII samples with Sr exchanged for Ca have shown two Ê [42]. It is believed that strontium has taken the MnASr distances of 3.5 A position normally occupied by calcium. Ca-depleted samples are inactive because S2Tyrz is the highest state that can be attained. Replacement of chloride by bromide [43] does not affect activity but iodide partially and ¯uoride completely inhibits O2 evolution. Fluoride binding appears to elongate Ê EXAFS Mn  Mn vector to 2.8 A Ê , suggesting binding at or near the the 2.7 A Mn cluster. Important information for making Mn oxidation state assignments for the S states comes from NMR proton relaxation studies (NMR-PRE) [44], where Mn(II) and Mn(IV) are strongly relaxing and Mn(III) weakly relaxing ions. Microwave power saturation studies [44b] of the S2 EPR signal of PSII can also probe the oxidation states involved. In conjunction with XANES data [40], both NMR and EPR results suggest that an Mn(II) exists in S0, that one Mn(III) is oxidized to Mn(IV) in the S1 to S2 conversion, and that no change in Mn oxidation state occurs on going from S2 to S3. An extra paramagnet produced in S3 and coupled to the Mn cluster is thought [44b] to prevent observation of the S2 multiline spectrum in S3. In the process when Tyrz is oxidized by P680 it loses a proton to become a neutral tyrosyl radical. In some way Tyrz is recreated by obtaining an electron and a proton in each step of water oxidation. Even though the details of this process are still under debate, this is one of the most important experimental ®ndings on which the models discussed below are built. Independently of the mechanism of this process, it means that the energy available to the water oxidizing complex in each step is approximately equal to the bond strength of the Tyrz OAH bond, which is equal to 86.5 kcal/mol. This quantity can be

Quantum Chemical Studies on Metal-Oxo Species

135

modi®ed, but only slightly, by changes of the charge of the cluster and changes in hydrogen bonding occurring during the S-state transitions. Two leading models for recreation of Tyrz exist. In the ®rst, termed the hydrogen abstraction mechanism, [45], see Fig. 3a, the tyrosyl radical obtains the proton and electron from a water molecule coordinated to manganese in a hydrogen atom transfer step. In the second model, the electron transfer model, see Fig. 3b, the tyrosyl radical obtains the electron from the manganese complex and the proton from a nearby protonated base. In that model, the water molecules which will eventually form water will lose their protons to a different base. This model has recently been elaborated further and some new aspects and modi®cations of the proton translocation mechanism have been introduced [46]. The models for water oxidation, discussed below, are independent of which of these schemes is the correct one.

Fig. 3a,b. Schematic picture of: a the hydrogen abstraction scheme; b the electron transfer

scheme

136

P.E.M. Siegbahn á R.H. Crabtree

Not only did Babcock et al. [45a±e] propose H atom abstraction from bound water as the key step in OAH bond breaking but they also suggested a role for metalloradicals in O2 formation via an Mn(IV)@O/(HO)Mn(IV) coupling induced by H transfer to Tyrz. Baldwin and Pecoraro [45f] have experimental evidence from model compounds that lAOAH bonds are relatively weak in Mn2(III,IV) dimers, consistent with H abstraction being important. Yachandra et al. have considered a pair of bridging oxyl radicals as precursors to OAO bond formation [36a]. Both Pecoraro et al. [45g] and Brudvig et al. [47b] have discussed OAO bond formation via nucleophilic attack by a hydroxo group on an Mn(V)@O oxo species. The proposed overall requirement for charge neutrality has also been emphasized by several authors as most appropriate for a cluster buried deep in the low dielectric medium of the protein [45, 48]. Krishtalik [49] applied a thermodynamic analysis to the problem and came to several conclusions about the O2-evolution mechanism. He proposed that the barrier produced by the repulsion of unbound O-atoms during their mutual approach may be partly surmounted using the binding energy of water molecules to the manganese ions. The concerted transfer of electrons and protons [45, 48] involving the participation of bases stronger than water was considered to improve the process energetics. The most likely reaction path was considered to be rate-determining two-electron water oxidation to the hydrogen peroxide oxidation level, followed by two fast oxidation steps of H2O2 to HO2 and on to O2. In 1992, Proserpio et al. [50] looked at possible O2 evolution mechanisms with the extended HuÈckel method. The oxygen-oxygen coupling process was examined starting from some known Mn cluster types and assuming oxo-oxo coupling as the route to O2. Tetranuclear models were also considered. An inplane approach of two oxo ligands was proposed to have the lowest barrier for peroxide bond formation. A slight energetic advantage was suggested in the peroxide to dioxygen step when the oxo ligands are coordinated to at least three Mn ions compared to coordination to only two Mn ions. No clear cut preferred mechanism could be devised, however. In other early work, Proserpio et al. [51] also used the extended HuÈckel method to study the problem. The energies of two complexes containing peroxo groups bound at a dinuclear fragment within a tetranuclear cluster (`dimer-of-dimers' model) were minimized using the universal force ®eld technique. The equilibrium geometries of the peroxo bound tetranuclear models indicated asymmetric coordination of the peroxo oxygen with MnAOAOAMn torsion angles of ca. 60°, consistent with ground state (triplet) oxygen release by PSII. A recent series of papers [48, 52±54] developed a quantum chemical picture of the photosynthetic O2 evolution problem. One paper [48a] examined the energetic requirements for breaking OAH bonds of water and hydroxide ligands coordinated to manganese. Another reported initial proposals on how the OAO bond might be formed [52]. In the third and fourth papers [53] ± the most relevant ones to this review ± the properties of the key oxo species that promotes OAO bond formation were discussed. In addition, a number of general principles emerged which strongly constrain possible pathways.

Quantum Chemical Studies on Metal-Oxo Species

137

In the ®rst paper [48a] the H abstraction model was tested via quantum chemical calculations. The B3LYP method was used, and both monomeric and dimeric manganese model systems were studied. It was found that, by coordination to a manganese center, the ®rst OAH bond strength of water is lowered to a value 0.2 kcal/mol lower than that in tyrosine ± a critical point as described above. The second hydrogen abstraction energy was quite similar. Since thermoneutrality in the reaction (or a weak exothermicity) is a requirement for the hydrogen abstraction model, these calculations are in accord with this model. In the second paper [52], several pathways for OAO bond formation were examined. For example, the approach of two Mn@O oxos to form O2 were investigated, as well as the approach of one MnAOH hydroxo towards an Mn@O oxo to form a peroxide. All of these attempts led to exceedingly high barriers for OAO bond formation. The reason for this could be traced to the fact that in all cases an oxygen radical, oxyl, is formed early along the reaction pathways, which for the systems investigated required too much energy. Further investigations of OAO bond formation pathways therefore focused on this particular problem (see below). The third paper [53] proposed a full mechanism based on a few quite general results and principles that were thought to rule out a very large number of otherwise viable pathways. One of these refers to the question of spin states. In redox reactions of a paramagnetic metal cluster, the starting and ®nal species may be on different potential energy surfaces and therefore have different ground spin states. In such a case there must be a transfer from the starting surface to the product surface at some point as one goes from starting material to products. In the case of the OEC Mn cluster, a reduction of the Mn ions occurs as the OAO bond is formed in one of the later S-states, which means that the local spin on manganese increases in this step. This leads directly to a spin state problem of the type mentioned. To take a speci®c case, if a 5-coordinate complex (H2O)(HO)3Mn(V)@O were to convert to (H2O)(HO)2Mn(III)AOAOAH in one step of O2 formation, the spin would rise from triplet in the starting material to quintet in the product; similar changes happen whatever the precise structures involved. In the OAO formation step, the reaction could occur on the potential energy surface of the reactants or on that of the products. In the former case, an excited state of the products will be formed; in the latter, the reaction has to start in an excited state for the reactants. There is in practice no third alternative as has been discussed, for example, for the simple reaction between an Ni atom and water [3] and also by Shaik et al. [1] who recently considered this point in the case of P-450. In the present case of OAO bond formation a large number of model calculations have shown that it is the excited state of the reactants that needs to be low-lying for obtaining a low barrier. Mn(V) oxo species with biologically relevant ligands have at least two spin states, a triplet with an Mn(V)@O structure and a quintet (or antiferromagnetic triplet) with an Mn(IV)AO· oxyl structure. The most stable structures of Mn(V) oxo are 5-coordinate while those of Mn(IV) oxyl are six coordinate. Six coordination stabilizes the Mn(IV) oxyl form relative to the Mn(V) oxo

138

P.E.M. Siegbahn á R.H. Crabtree

because Mn(IV), as a d3 ion, has a greater preference for an octahedral environment than does d2 Mn(V). As might be anticipated, the Mn(IV) oxyl, having an open shell oxo and an MnAO single bond, is far more reactive and interest centered on factors that favor formation of this state. Five coordinate Mn(V) oxo species like those that have been isolated and characterized tend to have quintet states at very high energy and so they are weakly reactive, consistent with their isolation as stable species. On the other hand, sixcoordinate species have accessible quintet states, consistent with the fact that none has ever been isolated. One experimental system [55] that is thought to involve a six-coordinate Mn(V) oxo species is found to evolve O2 catalytically, in line with the theoretical expectation. It should furthermore be added that, for complexes where the quintet state is low lying, the triplet state is also found to have oxyl radical character. A serious problem with this approach is that, given the OH bond strength of TyrOAH, only unreactive oxo complexes can be directly obtained via H abstraction from MnAOH by TyrO·, the H abstractor present in the OEC, via the route of Eqs. (1) and (2): Mn(II)…HOAH† ‡ TyrO ˆ Mn(III)OAH ‡ TyrOAH

…1†

Mn(III)OAH ‡ TyrO ˆ Mn(IV)@O ‡ TyrOAH

…2†

Reactive oxo complexes have very strong OAH bonds in the precursors, typically 20 kcal/mol larger than that in tyrosine. No other hydrogen abstraction pathways leading to the same 6-coordinated oxyl complex proved to be possible. The solution ®nally proposed is represented by Fig. 4. Only if one of the four Mn ions alone changes oxidation state can a feasible route to the desired oxyl be constructed. The initially 5-coordinate MnA is Mn(II) and bears an H2O ligand. Two H abstractions give an unreactive 5-coordinate Mn@O at the S2 state. At this point conversion to 6-coordination was proposed via chelation of the Ca cofactor via a bridging water. A third H abstraction is now possible from a bridging OH2 to give an oxo or oxyl at S3. Whether the oxo or oxyl is the ground state is not important ± it is only necessary that the oxyl state be accessible. In the key step, S3¢¢ ® S3¢¢¢, OAO bond formation takes place by a reaction between the oxyl and an outer sphere OH2 molecule (Fig. 5). In this step an H atom is transferred from the H2O to a l-oxo group to form an incipient outer sphere OH radical that can readily form an OAO bond. The resulting Mn(III)AOAOAH species readily loses a fourth H atom to give O2. The assignment of the S0, S1, and S2 states are consistent with the main experimental information available, as con®rmed by the similarity of the oxidation states suggested with models previously proposed by Klein et al. [36, 40] and by Hoganson and Babcock [45e] who both relied purely on experimental information. In particular, the NMR-PRE [44a], EPR power saturation [44b], and the XANES [40] data indicate Mn cluster oxidation state changes with S state, entirely in accord with all three proposals.

Quantum Chemical Studies on Metal-Oxo Species

139

Fig. 4. Proposed sequence of S-states using the dimer-of-dimers model for the OEC

The model is also in reasonable accord with the Messinger 18O labeling studies [41] because one O of O2 comes from the slow-exchanging oxo and the other from the fast-exchanging second coordination sphere water. The paramagnet interacting with the Mn cluster in S3, proposed in early work [44], is identi®ed here as the oxyl center. Finally, the S2 to S3 step has the

140

P.E.M. Siegbahn á R.H. Crabtree

Fig. 5. Structure of the transition state for OAO bond formation using a simple model with

one manganese center

highest activation energy, consistent with the structural rearrangement postulated here. Although the suggested oxygen radical mechanism solves many of the more signi®cant problems in water oxidation, there are still some questions that need more investigation. For example, in the model there is only one redoxactive manganese atom, and the function of the other three Mn-centers therefore remains to be found. Also, the suggested mechanism requires the formation of a terminal Mn(IV)@O oxo-group at S2, while continued model studies have shown that this is far from a trivial point. The previous model calculations furthermore only considered the most critical steps of water oxidation in detail, so detailed molecular models are still not suggested for the other steps. Finally, although the previous model gives a nice rationalization for why one oxygen of O2 rapidly exchanges with solvent water, it is not clear how a terminal Mn(IV)@O oxygen could exchange even slowly with solvent water which is found experimentally. In order to address some of these issues a large number of additional model calculations using a larger manganese cluster has recently been performed [54]. In most of these models there are three Mn-centers ± see Fig. 6 ± so the role of the fourth Mn has still not been addressed. The main conclusions from the Mn3-model calculations are as follows. The most essential features of the previous model remain the same, including the surprising fact that only one manganese is redox-active and that an oxyl-radical appears in S3. However, in the Mn3-model the oxyl does not appear as a terminal oxyl but as a l-oxo bridging oxyl. This means that the problem of creating a terminal oxo ligand, which is energetically very costly, is removed. The roles of the non-redox

Quantum Chemical Studies on Metal-Oxo Species

141

Fig. 6. The Mn3 model for the S2-state of the OEC

active Mn-centers are to tune the quite important and sensitive trans-effects that affect the OAH bond strengths. A list of six different important transeffects were found, including one which suggests an explanation for the surprising EXAFS-®nding that two of the Mn-Mn distances increase in the S2 to S3 transition. To achieve this structural change a cluster of the type shown in Fig. 6 is required, where the two l-oxo-bridging Mn-dimers are directly linked by the central Mn-atom in a bent con®guration. The role of the chloride cofactor is also to tune the trans-effects, while the function of the Ca-cofactor is to provide chelation energy, just as in the original oxyl-radical mechanism. To obtain maximum energetic effect by the chelation, Ca should bridge two of the three Mn-centers, leading to two short Mn-Ca distances in agreement with conclusions drawn from experiments for Sr-substituted clusters [42]. The water-oxidation is thought to occur by removing protons from water positioned in the corners of an incomplete cube where two Mn, one Ca, two l-oxo, and two waters form seven corners of the cube. The external water enters in the eighth, originally empty, corner. As in the previous model, this external water should easily exchange with solvent water. A mechanism for the slowly exchanging oxygen was also proposed, in which the bridging oxyl radical ®rst obtains a proton from an Mn-Ca bridging hydroxyl in a lowbarrier, water assisted process.

4 Conclusions The application of theoretical methods to the problem of enzyme mechanisms is in a very early stage of development and it is not yet certain how well our simple quantum models represent the enzyme. Benchmarking of the quantum chemical methods employed, however, suggests that inadequacies in these methods are not the main factors limiting the quality of the ®nal results. The procedure adopted in the work described above is to choose suf®ciently simpli®ed quantum model compounds to make the problem tractable and see whether the results obtained have promise. The results indicate that there are indeed suf®cient points of contact between the theoretical results obtained and the known biophysical and bioinorganic data to provide a solid basis for

142

P.E.M. Siegbahn á R.H. Crabtree

further work along these lines. In future work, more re®ned model calculations will be done to prove or disprove the ®ndings obtained so far. For example, once better structural data is obtained on the OEC, better quantum models will automatically become available. As described above, theoretical studies have already been very valuable in providing very clear energetic constraints to rule out otherwise plausible mechanisms that invoke energetically unreasonable structures or pathways. Indeed, few mechanisms survive this test. The one or ones that remain seem to be in good accord with the current experimental data and they also make predictions that have yet to be tested experimentally. As in all mechanistic work, other mechanisms cannot be ruled out. Nevertheless, both for MMO and PSII, signi®cant progress has already been made and we expect quantum chemical studies to be a growing presence in bioinorganic chemistry in the years ahead.

5 References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25.

Shaik S, Filatov M, Schroeder D, Schwartz H (1998) Chem Eur J 4: 193 Siegbahn PEM, Blomberg MRA, Svensson M (1993) J Am Chem Soc 115: 4191±4200 Mitchell SA, Blitz MA, Siegbahn PEM, Svensson M (1994) J Chem Phys 100: 423±433 Carroll JJ, Weisshaar JC, Siegbahn PEM, Wittborn CAM, Blomberg MRA (1995) J Phys Chem 99: 14,388±14,396 Mitchell SA (1992) In: Fontijn A (ed) Gas phase metal reactions. Elsevier, Amsterdam Bengali AA, Bergman RG, Moore CB (1995) J Am Chem Soc 117: 3879 Siegbahn PEM (1996) J Am Chem Soc 118: 1487±1496 (a) Shaik S, Danovich D, Fiedler A, SchroÈder D, Schwarz H (1996) Helvetica Chim Acta 78: 1393±1407; (b) Filatov M, Shaik S (1998) J Phys Chem A 102: 3835±3846 (a) Yoshizawa K, Shiota Y, Yamabe T (1998) Organometallics 17: 2825±2831; (b) Yoshizawa K, Ohta T, Shiota Y, Yamabe T (1997) Chem Lett 1213±1214; (c) Yoshizawa K, Ohta T, Shiota Y, Yamabe T (1998) Bull Chem Soc Jpn 71: 1899 Becke AD (1988) Phys Rev A 38: 3098 Becke AD (1992) J Chem Phys 96: 2155±2160 Becke AD (1993) J Chem Phys 98: 5648±5652 Siegbahn PEM, Blomberg MRA (1999) Ann Rev Phys Chem 50: 221±249 Siegbahn PEM, Blomberg MRA (1999) Chem Rev (in press) Feig AL, Lippard SJ (1994) Chem Rev 94: 759 Wallar BJ, Lipscomb JD (1996) Chem Rev 96: 2625 Dalton H (1980) Adv Appl Microbiol 6: 1 (a) SjoÈberg BM (1994) Structure 2: 793±796; (b) SjoÈberg BM (1997) Structure and Bonding 88: 139±173 (a) Stubbe J (1990) Biol Chem 265: 5330; (b) Mao SS, Holler TP, Yu GX, Bollinger JM, Booker S, Johnston MI, Stubbe J (1992) Biochemistry 31: 9733±9743 Ortiz de Montellano PR (1986) Cytochrome P-450: Structure, mechanism, biochemistry. Plenum Press, NY Wilkins PC, Wilkins RG (1987) Coord Chem Rev 79: 195 Rosenzweig AC, Frederick CA, Lippard SJ, Nordlund P (1993) Nature 366: 537 Andersson KK, Froland A, Lee SK, Lipscomb JD (1991) New J Chem 15: 411 (a) Stassinopoulos A, Caradonna J (1990) J Am Chem Soc 112: 7071; (b) Liesing RA, Brennan BA, Que L Jr (1991) J Am Chem Soc 113: 3988 Priestley ND, Floss HG, Frohland WA, Lipscomb JD, Williams PG, Morimoto H (1992) J Am Chem Soc 114: 7561±7562

Quantum Chemical Studies on Metal-Oxo Species

143

26. Liu KE, Johnson CC, Newcomb M, Lippard SJ(1993) J Am Chem Soc 115: 939±947 27. Choi SY, Eaton PE, Hollenberg PF, Liu KE, Lippard SJ, CC, Newcomb M, Putt DA, Upadhyaya SP, Xiong Y (1996) J Am Chem Soc 118: 6547±6555 28. Valentine AM, Lippard SJ (1997) J Chem Soc Dalton Trans 3925 29. Yoshizawa K, Yamabe T, Hoffmann R (1997) New J Chem 21: 151 30. Shestakov AF, Shilov AE (1996) J Mol Catal A 105: 1 31. Siegbahn PEM, Crabtree RH (1997) J Am Chem Soc 119: 3103 32. Shu L, Nesheim JC, Kauffmann K, Munck E, Lipscomb JD, Que L Jr (1997) Science 275: 515 33. Basch H, Mogi K, Musaev DG, Morokuma K (1999) J Am Chem Soc 121: 7249±7256 34. Siegbahn PEM (1999) J Inorg Chem 38: 2880±2889 35. Shaik S, private communication 36. (a) Yachandra VK, Sauer K, Klein MP (1996) Chem Rev 96: 2927±2950; (b) Ruettinger W, Dismukes GC (1997) Chem Rev 97; (c) Brudvig GW (1990) In: Hoff AJ (ed) Advanced EPR: applications in biology and biochemistry. Elsevier, Amsterdam, pp 839±865; (d) Yocum CF (1992) In: Pecoraro VL (ed) Manganese redox enzymes. VCH, NY, pp 71±83; (e) Kambara T, Govindjee (1985) Proc Nat Acad Sci USA, 82: 6119; (f) Pecoraro VL (1988) Photochem Photobiol 48: 249; (g) Christou G (1989) Acc Chem Res 22: 328; (h) Armstrong WH (1992) In: Pecoraro VL (ed) Manganese redox enzymes. VCH, NY, pp 261±286; (i) Yocum CF (1991) Biochim Biophys Acta 1059: 1 37 Lindberg K, Andreasson L (1992) Biochemistry 35: 14,259 38. Kok B, Forbush B, McGloin M (1970) Photochem Photobiol 11: 457 39. (a) Hansson O, Aasa R, Vanngard T (1987) Biophys J 51: 825; (b) de Paula JC, Brudvig GW (1985) J Am Chem Soc 107: 2643 40. (a) Roeloffs TA, Liang W, Latimer MJ, Cinco RM, Rompel A, Andrews JC, Sauer K, Yachandra VK, Klein MP (1996) Proc Natl Acad Sci USA 93: 3335; (b) Ono T, Noguchi T, Inoue Y, Kusinoki M, Matsushita T, Onayagi H (1992) Science 258: 1335; (c) Iuzzolini L, Dittmer J, DoÈrner W, Meyer-Klaucke W, Dau H (1998) Biochemistry 37: 17,112±17,119 41. (a) Messinger J, Badger M, Wydrzinski T (1995) Proc Natl Acad Sci USA 92: 3209±3213; (b) Messinger J, Hillier M, Badger M, Wydrzinski T (1995) In: Mathis P (ed) Photosynthesis: from light to biosphere. Kluwer, Dordrecht, vol 2, pp 283 42. Cinco RM, Robblee JH, Rompel A, Fernandez C, Yachandra VK, Sauer K, Klein MP (1998) J Phys Chem B102: 8248±8256 43. Beck WF, Brudvig GW (1988) Chem Scr 28A: 93 44. (a) Sharp RR (1992) In: Pecoraro VL (ed) Manganese redox enzymes. VCH, NY, pp 177± 196; (b) Styring SA, Rutherford AW(1988) Biochemistry 27: 4915; (c) Evelo RG, Styring SA, Rutherford AW, Hoff AJ (1988) Biochim Biophys Acta 973: 428 45. (a) Hoganson CW, Lydakis-Simantiris N, Tang XS, Tommos C, Warncke K, Babcock GT, Diner BA, McCracken J, Styring S (1988) Photosynth Res 46: 177; (b) Babcock GT (1995) In: Mathis P (ed) Photosynthesis from light to biosphere. Kluwer, Dordrecht, vol 2, pp 209; (c) Tommos C, Tang XS, Warncke K, Hoganson CW, Styring S, McCracken J, Diner BA, Babcock GT (1995) J Am Chem Soc 117: 10,325; (d) Tommos C, Babcock GT (1998) Acc Chem Res 31: 18; (e) Hoganson CW, Babcock GT (1997) Science 277: 1953; (f) Baldwin MJ, Pecoraro VL (1996) J Am Chem Soc 118: 11,325; (g) Pecoraro VL, Baldwin MJ, Caudle MT, Hsieh WY, Law NA (1998) Pure Appl Chem 70: 925; (h) Caudle MT, Pecoraro VL (1997) J Am Chem Soc 119: 3415 46. (a) Haumann M, BoÈgershausen O, Cherepanov D, Ahlbrink R, Junge W (1997) Photosynth Res 51: 193±208; (b) Ahlbrink R, Haumann M, Cherepanov D, BoÈgershausen O, Mulkidjanian A, Junge W (1998) Biochemistry 37: 1131±1142; (c) Haumann M, Junge W (1999) Biochim Biophys Acta 1411: 86±91 47. (a) Limburg J, Crabtree RH, Brudvig (1998) In: Meunier B (ed) Biomimetic oxidations catalyzed by transition metal complexes. ICP Press, Singapore; (b) Limburg J, Szalai VA, Brudvig GW (1999) J Chem Soc Dalton 1353

144

P.E.M. Siegbahn á R.H. Crabtree

Ê kermark B, Korall P (1997) 48. (a) Blomberg MRA, Siegbahn PEM, Styring S, Babcock GT, A J Am Chem Soc 119: 8285±8292; (b) Blomberg MRA, Siegbahn PEM (1997) Theor Chem Acc 97: 72±80 49. (a) Krishtalik LI (1986) Biochim Biophys Acta 849: 162; (b) Krishtalik LI (1990) Bioelectrochem Bioenerg 23: 249 50. Proserpio DM, Hoffmann R, Dismukes GC (1992) J Am Chem Soc 114: 4374 51. Proserpio DM, Rappe AK, Gorun SM (1993) Inorg Chim Acta 213: 319 52. Siegbahn PEM (1997) In: Comba P, Banci L (eds) Molecular modeling and dynamics of bioinorganic systems. Kluwer, pp 233±253 53. Siegbahn PEM, Crabtree RH (1999) J Am Chem Soc 121: 117±127 54. Siegbahn PEM Inorg Chem (submitted) 55. Limburg J, Vrettos JS, Liable-Sands LM, Rheingold AL, Crabtree RH, Brudvig GW (1999) Science 283: 1524

Characterization and Properties of Non-Heme Iron Peroxo Complexes Jean-Jacques Girerd1, FreÂdeÂric Banse2, Ariane Jalila Simaan3 Laboratoire de Chimie Inorganique, UMR CNRS 8613, Universite Paris-Sud, 91405 Orsay cedex, France 1 E-mail: [email protected] 2 E-mail: [email protected] 3 E-mail: [email protected] Iron-peroxo Fe(III)O2 and hydroperoxo Fe(III)OOH systems are important intermediates between the initial Fe(II)-dioxygen adduct and the ``activated'' form of the catalytic site in many mono-iron biomolecules. To the same peroxidic level correspond, in diiron enzymes, bridged peroxo Fe(III)AOAOAFe(III) intermediates. This review is concerned with the preparation and spectroscopic characterization of such intermediates in non-heme chemical systems, the properties of the natural systems being quoted as references. Although none have been crystallized, it seems very likely that Fe(III)OOH systems present a g1coordination mode for the hydroperoxo group. These Fe(III)OOH units have given clear signatures in UV-vis, resonance Raman and mass spectrometry. By EPR it was found that in Fe(III)OOH, the Fe(III) is low-spin (S = 1/2) and we propose here a simple rationalization of the characteristics of the EPR g-tensor. The electronic properties of the Fe(III)(g1-OOH) known so far, point toward a strong FeAO bond and a weak OAO bond, in total agreement with the reactivity scheme implying a cleavage of the OAO bond to lead formally to a Fe(V)O unit. Alkylperoxo systems are also included in this review. Fe(III)-peroxo systems Fe(III)O2 have been prepared and described. They contain high-spin Fe(III) and those identi®ed seem to be of the g2 type. The FeAO bond is weaker and the OAO one is stronger than in the Fe(III)OOH systems. The implication of these Fe(III)O2 units in catalysis is unclear. ``Complementary'' systems, such as Fe(III)(g1-OO) or Fe(III)(g2-OOH) have been evoked in publications but not identi®ed spectroscopically. These systems certainly deserve to be actively looked for. Finally, bridged peroxo Fe(III)AOAOAFe(III) systems have been characterized and most remarkably even crystallized and studied by X-ray diffraction in three cases. Their con®guration was found cis-planar or cis-gauche. Here the main question is how enzymes and possibly models can go from the Fe(III)AOAOAFe(III) state to the dil-oxo Fe(IV)O2Fe(IV) active state? Keywords: Peroxo complexes, Iron, Non-Heme, EPR, Resonance Raman, UV-vis

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

2

Fe-Hydroperoxo Units Fe(III)OOH . . . . . . . . . . . . . . . . . . . . . 149

2.1 2.2 2.3 2.4

UV-Visible Spectroscopy . . . . . . Mass Spectrometry . . . . . . . . . . EPR Spectroscopy . . . . . . . . . . . Resonance Raman Spectroscopy .

3

Fe-Alkylperoxo Units Fe(III)OOR . . . . . . . . . . . . . . . . . . . . . . 159

3.1 3.2

UV-Visible Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 EPR Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

149 151 151 156

146

J.-J. Girerd á F. Banse á A.J. Simaan

3.3 3.4 3.5

Resonance Raman Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . 161 Mass Spectrometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 Reactivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

4

Fe-Peroxo Units Fe(III)O2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

4.1 4.2 4.3 4.4

UV-Visible Spectroscopy . . . . . . EPR Spectroscopy . . . . . . . . . . . Resonance Raman Spectroscopy . Mass Spectrometry . . . . . . . . . .

5

Peroxo-Bridged Diiron Units Fe(III)OOFe(III) . . . . . . . . . . . . . 169

5.1 5.2 5.3

X-Ray Diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 UV-Visible Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 Resonance Raman Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . 171

6

Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

7

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

165 166 166 169

1 Introduction Dioxygen interaction with a heme-iron is of fundamental importance in several proteins implied in aerobic life processes. Considering for instance human heme proteins, the following examples can be quoted: 1. Dioxygen transport proteins like myoglobin or hemoglobin [1] 2. The cytochrome P450 which is so important in the metabolism of exogenous molecules, among them medicaments [2] 3. The cytochrome c oxidase, the terminal respiratory enzyme able to reduce dioxygen into water [3]. In myoglobin or hemoglobin [1] the dioxygen molecule ligates to the Fe(II) ion in a reversible way: when the dioxygen pressure decreases, the FeAO bond is broken and the dioxygen molecule leaves the iron ion. In cytochrome P450, the ®xation of dioxygen is followed by reduction by one external electron which weakens the OAO bond. Double protonation induces the breakage of the OAO bond and one oxygen atom stays on the iron ion and the other is eliminated as a water molecule. The Fe(V)O unit or its equivalent PáFe(IV)O inserts its oxygen into a CAH bond to create a CAOH alcohol group (Scheme 1). In the cytochrome c oxidase, a dioxygen molecule is inserted in between an Fe(II) ion and a Cu(I) ion and is likely to form an Fe(III)AOAOACu(II) unit. This unit appears to undergo a reduction by two electrons and protonation to give two water molecules (Scheme 2). However, the real mechanism is more sophisticated [3].

Characterization and Properties of Non-Heme Iron Peroxo Complexes

147

Scheme 1.

Scheme 2.

Several other heme proteins interacting with dioxygen are found in nature. This fascinating type of reaction has been reviewed in other chapters of this book. In recent years, several biomolecules implying reaction of dioxygen with non-heme iron have been discovered. These systems are in fact numerous and it is likely that many remain to be discovered. Nature often uses different means to reach the same chemical result: some of these non-heme enzymes achieve, by a different route, biochemical functions that are otherwise performed by a heme system. For instance, dioxygen transport is effected in sea worms not by a myoglobin type protein but by hemerythrin, a diiron non-heme system [1]. One of the two Fe(II) ions is pentacoordinated and thus has an open site for the (reversible) coordination of the dioxygen. Dioxygen links to this Fe(II) ion to form an Fe(III)Fe(III)OOH adduct (Scheme 3). Methane monooxygenase [4] (MMO), found in some bacteria, has some analogy with cytochrome P450 in its ability to insert an oxygen atom into a CAH bond. It is so powerful that it can do so even on the poorly reactive methane molecule CH4, giving methanol CH3OH. Unlike cytochrome P450,

Scheme 3.

148

J.-J. Girerd á F. Banse á A.J. Simaan

MMO contains two iron ions at the active site, each one in a non-heme environment. The mechanism of the catalytic reaction is not entirely understood but the main features are now known (Scheme 4). Dioxygen gives a peroxo bridge between the two Fe(III) ions. One possibility is that this peroxo unit rearranges to give a highly oxidizing Fe(IV)Fe(IV) unit (see the chapter by Siegbahn and Crabtree in this book). Some analogy may also exist between Cytochrome P450 and Bleomycin (BLM) [5] which is a natural antibiotic synthesized by fungi to protect themselves against bacteria. It contains a non-heme iron ion which interacts with dioxygen to lead to ``activated bleomycin'' which has been shown to contain an Fe(III)OOH unit [6] which is thought to decompose to an Fe(V)O group (Scheme 5) able to initiate DNA cleavage. It has been discovered that bleomycin can ef®ciently cleave the DNA molecules of some tumor cells and it is thus used as a drug in chemotherapy of some human cancers (see the chapter by Burger in this book). Important research efforts have been directed toward the comprehension of the interaction between dioxygen and non-heme iron ions. Structural motifs implied are, as mentioned above: 1. The Fe(III)AOAOAFe(III) one formed upon insertion of dioxygen between two Fe(II) ions as in MMO 2. The Fe(III)OOH one formed when dioxygen reacts with an Fe(II) ion to give an adduct which is reduced by one electron and protonated, as in bleomycin and hemerythrin It is clear that in these Fe(III)AOAOAFe(III) and Fe(III)OOH units the dioxygen molecule is reduced to the peroxo level. These entities can indeed in some cases be prepared by reaction of Fe(III) (or Fe(II)) with hydrogen peroxide. Moreover, if an alkyl peroxide ROOH is used in place of HOOH, an Fe(III)-alkylperoxo can be prepared. It has also been shown that FeO2 systems can be prepared by deprotonation of Fe(III)OOH. Overall, the four types of entities known today with non-heme iron are depicted in Scheme 6.

Scheme 4.

Scheme 5.

Characterization and Properties of Non-Heme Iron Peroxo Complexes

149

Scheme 6.

This review is devoted to the chemical and structural description of these species and their spectroscopic signatures together with an elementary discussion of their electronic properties.

2 Fe-Hydroperoxo Units Fe(III)OOH Except for the ligand Ph4DBA2) which gives dinuclear iron complexes (see below) [7], all the ligands represented in Fig. 1 have allowed the preparation of mononuclear Fe-hydroperoxo Fe(III)OOH complexes. In general these complexes have been obtained by reaction of the Fe(II) complex with an excess of H2O2, following the reaction: excess H2 O2

Fe(II) ƒƒƒƒƒƒ! Fe(III)OOH

…1†

2.1 UV-Visible Spectroscopy

The mononuclear Fe(III)OOH complexes are characterized by an absorption in the visible region attributed to a LMCT from the hydroperoxo group to the Fe(III) ion (Table 1). This absorption gives these complexes a purple color. Apparently, in MeOH, for the series trispicen [8], trispicMeen [9, 10], TPEN [8], ettpen [11], N4py [12±14], and Py5 [15], the higher the number of pyridines coordinated to Fe(III) the lower the band is in energy: 1. With 3 pyridines (trispicen, trispicMeen, TPEN, ettpen) 531±541 nm 2. With 4 pyridines (N4py) 548 nm 3. With 5 pyridines (Py5) 592 nm. This can be fortuitous or related to the p-accepting power of pyridine which will increase that of Fe(III) and thus displace the LMCT toward low energies.

150

J.-J. Girerd á F. Banse á A.J. Simaan

Fig. 1. Ligands which have allowed the characterization of Fe(III)OOH complexes, named

according to literature

Complexes of PMAH [16] and Me,Me,m-xyl ligands [17] with Fe(II) have been reacted with O2 and a reducing equivalent to give the Fe(III)OOH unit: O2 ;H‡ ;e

Fe(II) ƒƒƒ! Fe(III)OOH

…2†

UV-vis informations for the LMCT hydroperoxo-Fe(III) band have not been reported with these ligands. With Ph4DBA2) a model of hemerythrin has been built by Mizoguchi and Lippard [7]. The Fe(II)Fe(II) complex reacts to give an Fe(III)Fe(III)OOH

151

Characterization and Properties of Non-Heme Iron Peroxo Complexes

Table 1. Wavelength (molar extinction coef®cient) of the hydroperoxo ® Fe(III) LMCT for

Fe(III)OOH units Complex

kmax …e=M

[(trispicMeen)Fe(OOH)]2+ [(trispicen)Fe(OOH)]2+ [(TPEN)Fe(OOH)]2+ [(ettpen)Fe(OOH)]2+ [(TPA)Fe(OOH)]2+ [(N4Py)Fe(OOH)]2+

537 531 541 536 538 532 548 592 470

[(Py5)Fe(OOH)]2+ ``[Fe2(l-Ph4DBA)(OOH)]''

nm nm nm nm nm nm nm nm nm

1

 cm 1 †

(1000) in methanol (950) in methanol (900) in methanol (1000) in methanol (1000) in acetonitrile (1100) in acetone (1100) in methanol (?) in methanol (2600) in CH2Cl2

Refs. [9, 10] [8] [8] [11] [18] [12±14] [15] [7]

hydroperoxo adduct which is a good mimic of the corresponding group in oxyhemerythrin. The LMCT hydroperoxo-Fe(III) band was observed at 470 nm. The higher energy is in agreement with a decrease in the electron af®nity of Fe(III) due to ligation to negative carboxylate groups. 2.2 Mass Spectrometry

Electrospray Ionization Mass Spectrometry (ESI-MS) appeared to be very useful in the identi®cation of the Fe(III)OOH intermediates. First, Sam et al. [6] demonstrated using ESI-MS, that activated bleomycin contains an Fe(III)OOH unit. The attribution was corroborated with 18O substitution. Several aminopyridine type model complexes have been similarly studied and have been proven to contain the same Fe(III)OOH unit [9±12, 15, 18]. Information on equilibria in solution has also been obtained as well as evidence for ligand degradation [9, 10] and Fe(IV)O species [11, 18]. The demonstration of the existence of the unit Fe(III)OOH in these species, added to considerations from the structure of the pentadentate ligands such as N4py or Py5, make very likely the attribution of an g1 coordination mode of the hydroperoxo group in these complexes (Scheme 7). 2.3 EPR Spectroscopy

Fe-hydroperoxo systems have been studied by EPR spectroscopy. EPR data are summarized in Table 2.

Scheme 7.

of the wave functions; S = a2 + b2 + c2; V is the rhombic distortion energy; D is the axial distortion energy; both are in unit of k, the spin-orbit coupling constant Complex

Ref.

±gmin (gz) gmax (gy)

±ginter (gx) a

b

c

S

V/k

D/k

(BLM)Fe(OOH) [(trispicMeen)Fe(OOH)]2+ [(trispicen)Fe(OOH)]2+ [(TPEN)Fe(OOH)]2+ [(bztpen)Fe(OOH)]2+ [(TPA)Fe(OOH)]2+ [(N4Py)Fe(OOH)]2+ [(Py5)Fe(OOH)]2+ [Fe(bpy)2(py)(OOH)]2+ [Fe(phen)2(py)(OOH)]2+ [(PMA)Fe(OOH)]+ ``[(Me,Me,m-xyl)Fe(OOH)]''

[19] [9, 10] [8] [8] [11] [18] [12] [15] [20] [20] [16] [17]

)1.94 )1.95 )1.96 )1.997 )1.96 )1.96 )1.98 )1.98 )1.97 )1.97 )1.94 )1.93

)2.17 )2.12 )2.14 )2.15 )2.16 )2.15 )2.12 )2.13 )2.11 )2.12 )2.17 )2.16

0.052 0.039 0.041 0.040 0.046 0.044 0.033 0.035 0.033 0.035 0.051 0.052

0.993 0.992 0.994 0.999 0.995 0.995 0.998 0.998 0.995 0.995 0.992 0.989

0.9930 0.9885 0.9930 1.0022 0.9952 0.9939 0.9989 0.9984 0.9929 0.9929 0.9900 0.9870

3.04 4.05 2.86 3.98 2.00 2.18 4.34 1.80 2.87 0.95 1.90 2.50

)8.74 )11.31 )11.19 )11.29 )10.45 )10.89 )13.79 )13.89 )14.31 )14.19 )9.14 )8.89

2.26 2.19 2.19 2.22 2.20 2.19 2.17 2.15 2.14 2.13 2.22 2.23

0.072 0.055 0.053 0.056 0.055 0.053 0.044 0.040 0.040 0.037 0.063 0.063

152

Table 2. EPR gmax, ginter, gmin values for Fe(III)OOH S = 1/2 units. Parameters of the Grif®th-Taylor model of the g tensor: a, b, c are the coef®cients

J.-J. Girerd á F. Banse á A.J. Simaan

Characterization and Properties of Non-Heme Iron Peroxo Complexes

153

In every case these signals arise from a low-spin S = 1/2 Fe(III). These spectra are closer to axiality than their Fe(III) precursors. This was also found with (BLM)FeIII (g = 2.45, 2.18, 1.89) and (BLM)FeIII(OOH) (g = 2.27, 2.17, 1.94) [19]. In Fig. 2 the g values are represented as a function of gmax. We observe that gmin slightly decreases and ginter slightly increases when gmax increases. The differences gmax ± gmin and gmax ± ginter increase when gmax increases. It is possible to describe the g values for an LS Fe(III) by using the model of Grif®th [21] as presented by Taylor [22]. This model takes into account spinorbit coupling and distortion from octahedral geometry in such low spin Fe(III) complexes. In octahedral geometry, Fe(III) has a 2T2g ground state. The effects mentioned above lift the sixfold degeneracy of the 2T2g state and three Kramers doublets are obtained. In general the degeneracy is lifted enough to give well-separated Kramers doublets and only that lowest in energy is substantially populated at 100 K. Taylor has given convenient expressions of the two wave functions of the ground Kramers doublet and of the associated g values. The results are presented in Table 2. As Veselov et al. [23] we made the hypothesis that gxgygz is positive. This hypothesis seems justi®ed by the chemical meaning of the results. As described by Taylor [22] we selected the axes in such a way that |V| < (2/3)|D|, where V is the rhombic distortion energy, D the axial distortion energy and k the spin-orbit coupling parameter that has been assumed constant for all these LS Fe(III) systems. For (BLM)FeIII(OOH), one ®nds gz = )1.94, gy = 2.27, gx = )2.17, a = 0.072,

Fig. 2. gmax (m), ginter (à), gmin (d) as a function of gmax for Fe(III)OOH S = 1/2 units: (a)

[Fe(phen)2(py)(OOH)]2+; (b) [Fe(bpy)2(py)(OOH)]2+; (c) [(Py5)Fe(OOH)]2+; (d) [(N4Py) Fe(OOH)]2+; (e) [(TPA)Fe(OOH)]2+; (f) [(trispicen)Fe(OOH)]2+; (g) [(trispicMeen) Fe(OOH)]2+; (h) [(bztpen)Fe(OOH)]2+; (i) [(TPEN)Fe(OOH)]2+; (j) [(PMA)Fe(OOH)]+; (k) ``[(Me,Me,m-xyl)Fe(OOH)]''; (l) (BLM)Fe(OOH)

154

J.-J. Girerd á F. Banse á A.J. Simaan

b = 0.051, c = 0.993, V/k = 3.04, D/k = )8.74 where the coef®cients a, b, and c de®ne the functions of the resonating Kramers doublet: j‡i ˆ ajyzai ibjxzai cjxybi j i ˆ ajyzbi ibjxzbi cjxyai

…3†

The values obtained mean that the hole (and hence the unpaired electron) is in an orbital which has a dominant dxy character, and that this orbital is strongly destabilized. The dxz and dyz orbitals would be lower in energy with the dxz the lowest (Scheme 8). As for model compounds, one can take the examples of [(PMA)FeIII(OOH)]+ with gz = )1.94, gy = 2.22, gx = )2.17, a = 0.063, b = 0.051, c = 0.992, V/k = 1.90, D/k = )9.14 and [(N4py)FeIII(OOH)]2+ with gz = )1.98, gy = 2.17, gx = )2.12, a = 0.044, b = 0.033, c = 0.998, V/k = 4.34, D/k = )13.79. These results show that [(PMA)FeIII(OOH)]+ and (BLM)FeIII(OOH) systems have similar tetragonal distortion energies ()9.14 and )8.74 in units of k) but that [(N4py)FeIII(OOH)]2+ has a much larger one ()13.70) in absolute value. This is apparent in Fig. 3 which presents a diagram of V/k as a function of )D/k. This diagram suggests that these compounds can be divided into three types: 1. Those which possess at least 4 pyridine groups and exhibit a large |D| 2. Those containing 3 pyridine groups which have an intermediate value of |D| 3. The complexes with BLM and PMAH which have a smaller |D| A possible explanation is that the more accepting the Fe(III) is, the stronger the interaction is with the donor p* antibonding orbital of the hydroperoxo group and thus the larger is |D|. This implies that the hydroperoxo is along x or y since the most destabilized orbital is the dxy. Veselov et al. [23] have studied the ENDOR of activated bleomycin and give an important suggestion for the orientation of the axes. They proposed, by

Scheme 8.

Characterization and Properties of Non-Heme Iron Peroxo Complexes

155

Fig. 3. Graphical representation of the Grif®th-Taylor model of the g tensor of Fe(III)OOH

S = 1/2 units. The rhombic distortion energy V is represented as a function of the opposite value of D, the axial distortion energy, both in units of k, the spin-orbit coupling constant. The letter code for the complexes is the same as in Fig. 2

analogy with low spin heme systems [24], that the axis of gmax (y from the calculation above) is perpendicular to the conjugated pyrimidine plane and parallel to the Fe-peroxo axis. This is not in contradiction with the discussion above and leads to the following proposal for the orientation of the g tensor (Scheme 9). In this scheme the g values reported are those of (BLM)FeIII(OOH) but it is very likely that this diagram is valid for the model complexes as well. EPR thus suggests that the transfer of electron density from the antibonding orbital of the hydroperoxo group must be more important for ligands of type 1

Scheme 9.

156

J.-J. Girerd á F. Banse á A.J. Simaan

than for type 2 which in turn will present a transfer more important than type 3. Thus the FeAO and OAO bonds must be stronger in the order 1 > 2 > 3, as far as the P effects are concerned. 2.4 Resonance Raman Spectroscopy

The last conclusion can be compared to the results obtained by Resonance Raman using an excitation into the LMCT band. Data are given in Table 3. The vibrations of a polyatomic species such as the FeOOH unit cannot be reduced to the stretchings of pairs of atoms. Nevertheless quite often a vibration includes a dominant contribution which is then used as a label. It is considered so in Table 3 but the approximate nature of such a description has to be remembered. To estimate the adequacy of the description of a vibration by the diatomic model, it is useful to compare the shift observed upon isotopic substitution (16O/18O) with that predicted with the diatomic vibrator OAO or FeAO formula p mF2 O ˆ 1302:8 fFeO …lFe ‡ lO † p mOO ˆ 1302:8 2fOO lO

…4†

Ê and li = 1/mi. where fi are the force constants in mdyn/A Several mononuclear cases have been studied and the results are pretty similar. With [(trispicMeen)FeIII(OOH)]2+ [8] two frequencies are observed at Table 3. Raman mFeO and mOO frequencies in cm)1 for Fe(III)OOH with

O16O or 18O18O. Ê ; fFeO = 2.93 Calculated values are noted with *: a) diatomic vibrators: fOO = 2.94 mdyn/A Ê ; b) diatomic vibrators: fOO = 2.99 mdyn/A Ê ; fFeO = 2.79 mdyn/A Ê ; c) diatomic mdyn/A Ê ; f6FeO = 2.86 mdyn/A Ê ; d) diatomic vibrators: fOO = 3.36 vibrators: fOO = 3.02 mdyn/A Ê ; fFeO = 1.86 mdyn/A Ê ; e) polyatomic vibrators [25a]: fOO = 3.29 mdyn/A Ê; mdyn/A Ê and other constants non zero fFeO = 1.91 mdyn/A 16

O16O

Complex

[(N4Py)Fe(OOH)]2+ 2+

[(TPA)Fe(OOH)] [(trispicMeen)Fe(OOH)]2+ [(TPEN)Fe(OOH)]2+ [(trispicen)Fe(OOH)]2+ Oxyhemet. ``[Fe2(l-Ph4DBA)(OOH)]''

18

16

O18O

Refs.

mOO

mFeO

mOO

mFeO

790 790*a 789 796 796*b 796 801 801*c 844 844*d 843*e 843

632 632* 626 617 617* 617 625 625* 503 503* 504* )

746 ()44) 745*()45) ± 752 ()44) 751*()45) ) 750 ()51) 756*()46) 796 ()48) 796*()48) 796*()47) 797 ()46)

616 ()16) 604*()28) ± 600 ()17) 590*()27) ) 602 ()23) 597*()27) 479 ()24) 481*()22) 477*()27) ±

[13, 14] [13] [8] [8] [8] [25a, 26, 27] [7]

Characterization and Properties of Non-Heme Iron Peroxo Complexes

157

796 cm)1 and 617 cm)1 which shift upon 16O/18O substitution to 752 cm)1()44) and 600 cm)1()17) respectively. With [(N4py)FeIII(OOH)]2+ [13, 14] two frequencies are observed at 790 cm)1 and 632 cm)1 which shift upon substitution to 746 cm)1()44) and 616 cm)1()16) respectively. The highest frequency shift is of the order expected for a pure OAO frequency but the shift of the lowest frequency is lower than expected for a pure FeAO vibration. The fact that the value of the FeAO frequency is larger for [(N4py)FeIII(OOH)]2+ than for [(trispicMeen)FeIII(OOH)]2+ could con®rm the analysis given in the EPR section above. However [(N4py)FeIII(OOH)]2+ presents a mOO frequency smaller than that of [(trispicMeen)FeIII(OOH)]2+ in apparent contradiction with the interpretation of the EPR. One has nevertheless to be cautious with such a qualitative model and more quantitative work is needed here. As we will see later, analogous frequencies have also been detected for other Fe/peroxo type systems. Figure 4 summarizes the data published for Fe(III)OOH, Fe(III)O2, and Fe(III)OOFe(III) units as mFeO = f(mOO). It is quite recognizable that the points in Fig. 4 form approximately one domain per type of unit. What is surprising in the experimental values for the Fe(III)OOH low spin systems is that the mFeO is the largest and mOO the smallest of all frequencies known for Fe(III)-peroxo systems (upper left part of Fig. 4). For instance, as shown in Table 3, the diatomic model implies similar fFeO and fOO force Ê ) which are quite different from those that have constants (around 2.9 mdyn/A been found for the Fe(III)OOH motif in hemerythrin by Brunold and Solomon Ê and fOO = 3.29 mdyn/A Ê ) [25a]. The reason for this is not (fFeO = 1.91 mdyn/A clear to us. An NCA analysis for the mononuclear Fe(III)OOH is certainly needed. Such a calculation could lead to improved force constants. The in¯uence of the FeOO angle has certainly to be studied as has the coupling between the stretching modes and the bending. We also found that H/D substitution has an effect on trispicen (801 cm)1 and 618 cm)1) [8] which points toward the need for an NCA extended to the four atoms of the FeOOH unit. However it is likely that the force constants will remain of the order of those found by the simple diatomic model. Interestingly, Roelfes et al. [14] have proposed that the unusually strong FeAO bond may be mainly related to the low spin nature of the Fe(III) ion. This is very convincing since in other systems of Fig. 4, in which the Fe(III) is high spin, the FeAO frequency is indeed much lower. In particular in the oxyhemerythrin motif Fe(III)OOH which contains a high spin Fe(III) the FeAO frequency is 503 cm)1 [25±27] and the representative point in Fig. 4 is clearly separated from the low spin Fe(III)OOH ones. Unfortunately Raman data have not been published for activated BLM. One can only propose from the model outlined above that in low spin activated BLM: 1. The FeAO bond is strong due to the LS state of Fe(III) 2. The OAO bond is weak since, due to the poor electron accepting power of BLM, the electron transfer from the p* antibonding peroxo orbital toward Fe(III) is weak

158

J.-J. Girerd á F. Banse á A.J. Simaan

Fig. 4. Graphical representation of the Raman mFeO and mOO frequencies (in cm)1) for

Fe(III)OOH (Ñ), Fe(III)O2 (s) and Fe(III)OOFe(III) (d) units. Note that each type of unit corresponds to one domain except the point corresponding to hemerythrin which is off the Fe(III)OOH domain (see text): (a) [(TPA)Fe(OOH)]2+; (b) [(N4Py)Fe(OOH)]2+; (c) [(TPEN)Fe(OOH)]2+; (d) [(trispicMeen)Fe(OOH)]2+; (e) [(trispicen)Fe(OOH)]2+; (f) oxyhemerythrin; (a0 ) [(EDTA)Fe(O2)]3); (b0 ) [(trispicMeen)Fe(O2)]+; (c0 ) [(TPEN)Fe(O2)]+; (a00 ) ferritin; (b00 ) [Fe2(l-O2)(l-O)(6-Me3TPA)2]2+; (c00 ) RNR mutant; (d00 ) Fe2(l-O2)(l-OBz) {HB(pz0 )3}2; (e00 ) [Fe2(l-O2)(Ph3PO)8 á 2H2O]4+; (f 00 ) [Fe2(l-O2)(l-O2CCH2Ph)2{HB (pz0 )3}2]; (g00 ) [Fe2(l-O2)(5,6-Me2±HPTB)(NO3)2]+; (h00 ) [Fe2(l-O2)(N-Et-HPTB)(NO3)2]+; (i00 ) [Fe2(l-O2)(HPTB)(NO3)2]+; (j00 ) [Fe2(l-O2)(N-Et-HPTB)(OBz)]2+; (k00 ) desaturase

3. The distal oxygen atom of the Fe(III)OOH group may be basic due to the negative ligands of Fe(III). These features would support models of BLM reactivity that imply an OAO bond breakage upon protonation to generate an oxidizing Fe(V)O unit. For the dimer system oxyhemet, the agreement between the observed values and those calculated by the simple diatomic model is striking. The frequency mOO is larger and the mFeO is smaller than in the mononuclear systems discussed above. In the function of hemerythrin, the FeAO bond is broken and the OAO bond conserved, a difference with BLM which could ultimately be related to that in the spin state of the Fe(III) linked to the hydroperoxo group. For more details on hemerythrin, the reader is referred to a very complete work by Brunold and Solomon [25]. Some of these model compounds have shown a DNA cleavage activity similar to that of BLM [10, 28]. Recently Chen and Que studied the catalytic hydroxylation of alkanes by [(bispicMe2en)FeIII(OOH)]2+ [29], an aminopyridine type system analogous to those discussed here. They propose that the reaction may involve the heterolytic breakage of the OAO bond of the Fe(III)(g1-OOH) to form an Fe(V)O unit [29], in strong analogy to what is proposed for BLM [5].

Characterization and Properties of Non-Heme Iron Peroxo Complexes

159

The same authors demonstrate the possibility to do cis-dihydroxylation by a related Fe(III)OOH system using the ligand 6-Me3TPA (see Fig. 5 for the ligand) [30]. To explain this very interesting reactivity the authors proposed that with this ligand a Fe(III)(g2-OOH) would form, the OAO bond being broken later in the catalytic cycle (Scheme 10). The authors did not give spectroscopic information concerning the Fe(III)(g2-OOH) intermediate which, if it indeed exists, would be a new motif extremely interesting to explore.

Scheme 10.

3 Fe-Alkylperoxo Units Fe(III)OOR Several systems presenting the Fe(III)OOR group have been reported, most commonly with R = tert-butyl. The ligands of Fig. 5 were used as well as some ligands of Fig. 1. 3.1 UV-Visible Spectroscopy

As with the Fe(III)OOH-containing complexes, the Fe(III)OOR systems present a peroxo/Fe(III) LMCT band around 600 nm (Table 4). To our

160

J.-J. Girerd á F. Banse á A.J. Simaan

Fig. 5. Ligands which have allowed the characterization of Fe(III)OOR complexes, named

according to literature

knowledge no rationalization of the characteristics of the LMCT band has been offered in this series. 3.2 EPR Spectroscopy

Some of the Fe(III)OOR have S = 1/2 as found by EPR (Table 5). The EPR signal of these species can be analyzed by the method of Taylor [22] (Table 5). Figure 6 presents the rhombic distortion energy V/k as a function of )D/k. Approximately the same trend as in Fig. 3 is found. The system with PMAH presents a small value of |D|, those with phen or bpy a large value, and the TPA ones intermediate values. Some phen systems are found in the same area as the TPA ones. We propose that basically the same explanation is valid here with a strengthening of the FeAO bond upon increase of the p-accepting power of the auxiliary ligand. Very interestingly, Wada et al. [31] and Que et al. [32±34] have shown that Fe(III)OOR systems can be high spin (S = 5/2) with auxiliary ligands which exert a weak enough ligand ®eld. The 6-MeTPA and 6-PhTPA ligands even give systems which present both spin states. In general the HS systems are found to be rhombic with g = 4.3 (Table 6). There are notable exceptions in those systems, recently reported by Wada et al. [31] which exhibit an almost axial anisotropy (E/D close to 0.07). The Fe(III) precursor [Fe(bppa)(t-BuCOO)]2+ of these last complexes has been crystallized. By X-ray diffraction it has been found that the Fe(III) ion is

161

Characterization and Properties of Non-Heme Iron Peroxo Complexes

Table 4. Wavelength (molar extinction coef®cient) of the alkylperoxo ® Fe(III) LMCT for

Fe(III)OOR units Complex

kmax …e=M

[Fe(TPA)(H2O)(OOt-Bu)]2+ [Fe(6-MeTPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me2TPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me3TPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me3TPA)(OBz)(OOt-Bu)]+ [Fe(6-Me3TPA)(OBz)(OOcumyl)]+ [Fe(6-PhTPA)(OOt-Bu)]2+ [Fe(BPPA)(OOt-Bu)]2+ [Fe(BPPA)(OOcumyl)]2+ [Fe(TPA)(HOCH2Ph)(OOt-Bu)]2+ [Fe(bpy)2(HOCH2Ph)(OOt-Bu)]2+ [Fe(bpy)2(HOCH2Ph)(OOcumyl)]2+

600 598 552 562 510 506 630 613 585 598 640 627

nm nm nm nm nm nm nm nm nm nm nm nm

1

 cm 1 †

(2200) (2000) (2000) (2000) (2300) (2300) (?) (2000) (2200) (2000) (1700) (1700)

in in in in in in

MeCN MeCN MeCN MeCN CH2Cl2 CH2Cl2

in in in in in

MeCN MeCN MeCN MeCN MeCN

References [33] [33] [33] [33] [32] [32] [34] [31] [31] [35] [36] [36]

heptacoordinated to the three pyridine nitrogen atoms, the tertiary amino nitrogen atom, two oxygen atoms from the pivalamido groups, and one oxygen atom from the t-BuCOO) anion. This is a rather unusual structure. 3.3 Resonance Raman Spectroscopy

Resonance Raman data have been published (Table 7). Zang et al. [33] have shown that the low-spin and high-spin Fe(III)OOR have different spectra. The corresponding modes of vibrations are of mixed nature, implying also C-O and C-C stretchings. 3.4 Mass Spectrometry

The ions [{Fe(TPA)(OOt-Bu)}(ClO4)]+ and [{Fe(TPA)(OOt-Bu)}(ClO4)3]) have been detected by ESI-MS, demonstrating that the intermediate formed by reaction of t-BuOOH on [Fe2O(TPA)2(H2O)2]4+ is a monomeric iron(III)alkylperoxo complex [35]. Several other alkylperoxo complexes have been identi®ed using this technique [31, 33]. 3.5 Reactivity

Lange et al. [34] have studied the reactivity of the complex of 6-PhTPA with t-BuOOH. In a previous publication the authors showed that an Fe(III)OOR group undergoes a homolytic splitting into Fe(IV)O and  OR [38]. The ligand

Parameters of the Grif®th-Taylor model of the g tensor: a, b, c are the coef®cients of the wave functions; S = a2 + b2 + c2; V is the rhombic distortion energy; D is the axial distortion energy; both are in unit of k the spin-orbit coupling constant Complex

References ±gmin (gz) gmax (gy)

±ginter (gx) a

b

c

S

V/k

D/k

[Fe(TPA)(H2O)(OOt-Bu)]2+ [Fe(6-MeTPA)(H2O)(OOt-Bu)]2+(*) [Fe(TPA)(HOCH2Ph)(OOt-Bu)]2+ [Fe(PMA)(OOt-Bu)]+ [Fe(bpy)2(HOCH2Ph)(OOt-Bu)]2+ [Fe(bpy)2(py)(OOt-Bu)]2+ [Fe(phen)2(py)(OOt-Bu)]2+ [Fe(phen)2(MeOH)(OOt-Bu)]2+ [Fe(6-PhTPA)(OOt-Bu)]2+(*)

[33] [33] [35] [37] [36] [20] [20] [20] [34]

)2.14 )2.12 )2.12 )2.18 )2.12 )2.12 )2.12 )2.19 )2.12

0.037 0.034 0.034 0.055 0.032 0.034 0.037 0.030 0.034

0.998 0.996 0.996 0.991 0.998 0.996 0.993 0.995 0.996

1.001 0.9966 0.9961 0.9919 0.9994 0.9961 0.9908 0.9934 0.9961

3.49 5.44 4.96 2.89 4.97 4.96 3.19 5.56 4.96

)12.39 )12.40 )12.57 )8.10 )13.55 )12.57 )12.39 )14.29 )12.57

)1.98 )1.97 )1.97 )1.93 )1.98 )1.97 )1.96 )1.97 )1.97

2.19 2.20 2.19 2.28 2.18 2.19 2.165 2.16 2.19

0.048 0.053 0.050 0.077 0.046 0.050 0.047 0.044 0.050

162

Table 5. EPR gmax, ginter, gmin values for Fe(III)OOR S = 1/2 units, (*) indicates that the compound presents both spin states (see also Table 6).

J.-J. Girerd á F. Banse á A.J. Simaan

163

Characterization and Properties of Non-Heme Iron Peroxo Complexes

Fig. 6. Graphical representation of the Grif®th-Taylor model of the g tensor of Fe(III)OOR

S = 1/2 units. The rhombic distortion energy V is represented as a function of the opposite value of D, the axial distortion energy, both in unit of k, the spin-orbit coupling constant: (a) [(PMA)Fe(OOt-Bu)]+; (b) [Fe(6-MeTPA)(H2O)(OOt-Bu)]2+; (c) [Fe(6-PhTPA)(H2O) (OOt-Bu)]2+; (d) [Fe(bpy)2(py)(OOt-Bu)]2+; (e) [Fe(TPA)(HOCH2Ph)(OOt-Bu)]2+; (f ) [Fe(TPA)(H2O)(OOt-Bu)]2+; (g) [Fe(phen)2(py)(OOt-Bu)]2+; (h) [Fe(bpy)2(py)(OOtBu)]2+; (i) [Fe(phen)2(MeOH)(OOt-Bu)]2+

6-PhTPA was built to test the insertion of the oxygen atom into the conveniently oriented aromatic CAH bond. Indeed the formation of the phenol group was observed (Scheme 11). Isotopic labeling experiments using RO18OH showed the oxygen atom of the phenol group was the 18O [34]. This is indeed a remarkable result.

Table 6. EPR data for high-spin S = 5/2 Fe(III)OOR complexes, (*) indicates that the

compound presents both spin states (see Table 5) Complex

EPR data

Refs.

[Fe(6-MeTPA)(H2O)(OOt-Bu)]2+(*) [Fe(6-PhTPA)(H2O)(OOt-Bu)]2+(*) [Fe(6-Me2TPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me3TPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me3TPA)(OBz)(OOt-Bu)]+ [Fe(6-Me3TPA)(OBz)(OOcumyl)]+ [Fe(BPPA)(OOt-Bu)]2+

4.3 4.3 4.3 4.3 4.3 4.3 7.58, (E/D 7.76, (E/D

[33] [34] [33] [33] [32] [32] [31]

[Fe(BPPA)(OOcumyl)]2+

5.81, 4.25, 1.82 = 0.067) 5.65, 4.20, 1.78 = 0.070)

[31]

Complex 2+

[Fe(TPA)(HOCH2Ph)(OOt-Bu)] [Fe(TPA)(H2O)(OOt-Bu)]2+ [Fe(6-MeTPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me2TPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me3TPA)(H2O)(OOt-Bu)]2+ [Fe(6-Me3TPA)(OBz)(OOt-Bu)]+ [Fe(6-Me3TPA)(OBz)(OOcumyl)]+ [Fe(BPPA)(OOt-Bu)]2+ [Fe(BPPA)(OOcumyl)]2+ [Fe(bpy)2(HOCH2Ph)(OOt-Bu)]2+ [Fe(bpy)2(HOCH2Ph)(OOcumyl)]2+ [Fe(6-PhTPA)(H2O)(OOt-Bu)]2+

164

Table 7. Raman frequencies in cm)1 for Fe(III)16O16OR or Fe(III)18O16OR units

FeA16O16OAR

FeA18O16OAR

Refs.

490, 490, 488, 470, 468, 464, 432, 469, 493, 678, 696, 502,

± 488, 638, 657, 672, 778 ± ± 463, 612, 829, 867 ± 432, 626, 804, 842 ± ± ± ± ±

[35] [33] [33] [33] [33] [32] [32] [31] [31] [36] [36] [34]

696, 796 (LS) 696, 796 (LS) 682, 790 (LS) 469, 648, 842, 878 (HS) 647, 843, 881 (HS) 637, 842, 877 (HS) 648, 844, 880 (HS) 554, 650, 880 (HS) 629, 838, 873 (HS) 639, 838, 878 (HS) 808 (LS) 805 (LS) 683 (LS) 469, 644, 841, 873 (HS)

J.-J. Girerd á F. Banse á A.J. Simaan

165

Characterization and Properties of Non-Heme Iron Peroxo Complexes

Scheme 11.

4 Fe-Peroxo Units Fe(III)O2 With the EDTA4) ligand, an interesting FeO2 species has been obtained [39± 41]. Analogous species were observed by deprotonation of the Fe(III)OOH species formed with the ligands trispicMeen [42], TPEN [8], and Rtpen [11] (see above). 4.1 UV-Visible Spectroscopy

UV-vis data are reported in Table 8. The LMCT band in the visible region for [(EDTA)Fe(O2)]3) occurs at 520 nm [39±41]. This complex has been studied in detail by Neese and Solomon [41]. With the aminopyridine ligands, a LMCT band is observed around 750 nm, at a lower energy than that of [(EDTA)Fe(O2)]3). This is likely to be related to the stronger p acidity of the aminopyridine type ligands. For these ligands the deprotonation of the Fe(III)OOH upon addition of a base was detected by the spectacular change in color from purple to blue as demonstrated by the UV-visible spectra of [(trispicMeen)Fe(OOH)]2+ and [(trispicMeen)Fe(O2)]+ (Fig. 7) [42]. 4.2 EPR Spectroscopy

The [(EDTA)Fe(O2)]3) complex has been found to be high spin rhombic (Table 9). Table 8. Wavelength (molar extinction coef®cient) of the peroxo ® Fe(III) LMCT for

Fe(III)O2 units

Complex

kmax …e=M

[(EDTA)Fe(g2-O2)]3) [(trispicMeen)Fe(g2-O2)]+ [(TPEN)Fe(g2-O2)]+ [(ettpen)Fe(g2-O2)]+

520 740 755 747

nm nm nm nm

1

 cm 1 †

(520) (500) (450) (500)

in in in in

water MeOH MeOH MeOH

Refs. [39, 40, 41] [42] [8] [11]

166

J.-J. Girerd á F. Banse á A.J. Simaan

Fig. 7. UV-vis spectra of [(trispicMeen)Fe(OOH)]2+ (dashed line) and [(trispicMeen)

Fe(O2)]+ (bold line) in MeOH (from Simaan et al. [42])

The ZFS parameters are found to be E/D = 0.21 and D = )1 ‹ 0.25 cm)1 [41]. The aminopyridine type systems have a different EPR signature: for instance, [(trispicMeen)Fe(O2)]+ [42] (Table 9) has g values which indicate E/D = 0.08. The D value has not been determined. Table 9. EPR data for high-spin S = 5/2 Fe(III)O2 complexes

Complex

EPR data

Refs.

[(EDTA)Fe(g2-O2)]3) [(trispicMeen)Fe(g2-O2)]+ [(TPEN)Fe(g2-O2)]+ [(bztpen)Fe(g2-O2)]+

E/D = 0.21 and D = )1 ‹ 0.25 cm)1 g = 7.5, 5.9 (E/D = 0.08) g = 10, 8.1, 5.6, 3.2, 2.4 (E/D = 0.1) g = 7.60, 5.74

[40, 41] [42] [8] [11]

4.3 Resonance Raman Spectroscopy

Raman spectra have been measured and proved the g2-coordination mode of the peroxo group (Table 10). The NCA treatment of FeO2 taking into account fOO for the OAO bond and fFeO for the FeAO bonds has been published [41]. The expressions given below use the angle OFeO = 2a as a parameter (Scheme 12). The number of frequencies is 3n ± 6 = 3. They are indicated below: p mFeOas ˆ 1302:8 ‰f FeO …lO ‡ lFe …1 cos 2a††Š q p m ˆ 1302:8 ‰…A  B†=2Š

…5†

167

Characterization and Properties of Non-Heme Iron Peroxo Complexes Table 10. Raman mOO, mFeO, and mFeOas frequencies in cm)1 for Fe(III)O2 with

16 16 O O or Ê; O18O. Calculated values are noted with *: a) diatomic vibrators: fOO = 3.14 mdyn/A Ê ; b) polyatomic vibrators calculation [41]: fOO = 3.02 mdyn/A Ê; fFeO = 1.55 mdyn/A Ê ; c) diatomic vibrators: fOO = 3.17 mdyn/A Ê ; fFeO = 1.61 mdyn/A Ê; fFeO = 1.55 mdyn/A Ê ; fFeO = 1.60 mdyn/A Ê d) polyatomic vibrators calculation: fOO = 3.015 mdyn/A

18

Complex

[(EDTA)Fe(g2-O2)]3)

16

O16O

O18O

Refs.

mOO

mFeO

mFeOas

mOO

mFeO

mFeOas

816

459

±

459*

±

820*b

462*

421*

468

±

820*c

469*

±

821*d

468*

427*

821

468

±

446 ()13) 439* ()20) 445* ()17) 454 ()14) 448* ()19) 451* ()17) ±

±

816*a

776 ()40) 770* ()46) 774* ()46) 776 ()44) 773* ()47) 775* ()46) ±

[(trispicMeen)Fe(g2-O2)]+ 820

[(TPEN)Fe(g2-O2)]+

18

[40, 41]

± )398* ()23) ±

[8]

± 405* ()22) ±

[8]

with A ˆ f FeO …lO ‡ lFe …1 ‡ cos 2a†† ‡ 2f OO lO B ˆ ‰f FeO …lO ‡ lFe …1 ‡ cos 2a††

2f OO lo Š2 ‡ 8f FeO f OO l2O sin2 a

One may notice that for a = 0 the vibrations can be considered as mFeOas, mFeOs, mOO. For a non-zero angle, some mixing occurs but this remains small if the angle is small, in such a way that m+ can be considered as mOO and m) as mFeOs. The resonance Raman data for [(EDTA)FeIII(O2)]3) have been reported as mOO = 816 cm)1 and mFeOs = 459 cm)1. Upon 16O/18O substitution, those Ê and frequencies shift to 776 cm)1 and 446 cm)1. Taking fOO = 3.02 mdyn/A )1 )1 Ê fFeO = 1.55 mdyn/A one obtains mOO = 820 cm and mFeOs = 462 cm for 16 16 O O and mOO = 774 cm)1 and mFeOs = 445 cm)1 for 18O18O [41]. The agreement is satisfying. The vibrations are rather ``pure'' since using the diatomic model one calculates the following shifted frequencies mOO = 770 cm)1 and mFeO = 439 cm)1 with 18O18O (see Table 10). Another example is [(trispicMeen)Fe(O2)]+. It was found that mOO = 820 cm)1 and mFeOs = 468 cm)1 and that upon 16O/18O substitution, Ê and they shift to 776 cm)1 and 456 cm)1 [8]. Taking fOO = 3.015 mdyn/A

Scheme 12.

168

J.-J. Girerd á F. Banse á A.J. Simaan

Ê one obtains mOO = 821 cm)1 and mFeOs = 468 cm)1 for fFeO = 1.6 mdyn/A 16 16 O O and mOO = 775 cm)1 and mFeOs = 451 cm)1 for 18O18O. This suggests that the FeAO bonds with the neutral pyridine ligand are stronger than with EDTA4) which makes sense if one considers the electron acceptor role of pyridine. One remarks that in Fig. 4, the FeO2 systems occupy a well de®ned zone which is different from the FeOOH and FeOOFe systems. In comparison with FeOOH, the weaker FeAO bond in FeO2 systems may be due to their HS state. Moreover the stronger OAO bond could imply a lower oxidative activity. Indeed it has been reported that [(EDTA)Fe(O2)]3) is unreactive toward organic substrates at pH > 10.5 while it becomes very reactive when the pH decreases [39]. These observations strongly suggest that the oxidant species is or arises from a protonated peroxo species. The geometry proposed for [(EDTA)Fe(O2)]3) is depicted in Scheme 13 [41] and for [(trispicMeen)Fe(O2)]+ in Scheme 14 [42]. That with the ligand trispicMeen is more tentative and one cannot dismiss the heptadentate structure. 4.4 Mass Spectrometry

Formation of the Fe(III)-peroxo complexes has been con®rmed by ESI-MS for the following aminopyridine systems: [(trispicMeen)Fe(O2)]+ (m/z 435) [42],

Scheme 13.

Scheme 14.

Characterization and Properties of Non-Heme Iron Peroxo Complexes

169

[(TPEN)Fe(O2)]+ (m/z 512.3) [8], and [(bztpen)Fe(O2)]+ (m/z 511.2) [11]. It has also been observed that this peak disappears as soon as the blue coloration disappears [8, 42].

5 Peroxo Bridged Diiron Units Fe(III)OOFe(III) The ligands allowing the preparation of Fe(III)AOAOAFe(III) motifs are represented in Fig. 8. They are of two types: 1. Non-assembling ligands such as HB(pz0 )3 or Mes2ArCO2 2. Assembling ligands which use either a phenolato or an alkoxo bridge 5.1 X-Ray Diffraction

In three cases a peroxo dimer has been crystallized and an X-ray diffraction study has revealed the structure [43±45]. In the complex Fe2(O2)(O2CCH2Ph)2{HB(pz0 )3}2 prepared by Kim and Lippard [43], the fragment FeOOFe has a cis-gauche conformation (dihedral angle = 53°, Fig. 9). Ê , 1.877(6) A Ê and the OAO distance is The FeAO distances are 1.881(6) A Ê . The two Fe(III) ions are bridged by two carboxylates and one 1.406(8) A peroxo dianion. Figure 10 reproduces the structure of [Fe2(O2)(Ph-bimp)(OBz)]2+ found by Ookubo et al. [44]. This complex uses the assembling heptadentate, monoanionic, ligand Ph-bimp). The two Fe(III) ions are bridged by the phenolato oxygen atom, one carboxylato group, and one peroxo. The fragment FeOOFe has a cis almost planar conformation (dihedral angle = 10°, Fig. 10). Ê , 1.864(4) A Ê and the OAO distance is The FeAO distances are 1.944(4) A Ê. 1.427(6) A Figure 11 gives the structure of [Fe2(O2)(N-Et-HPTB)(Ph3PO)2]3+ [45]. The assembling ligand is an heptadentate monoanion. The two Fe(III) ions are only doubly bridged here, by a peroxo and an alkoxo group. This imposes on the FeOOFe fragment a cis planar conformation (dihedral angle = 0°). The FeAO Ê and the OAO distance is 1.416(7) A Ê. distances are 1.880(4) A 5.2 UV-Visible Spectroscopy

UV-vis data are collected in Table 11 for some of the compounds (see Table 11 for references). The most thorough optical study has been published by Brunold et al. for Fe2(O2)(OBz)2{HB(pz0 )3}2 [46]. The main features related to peroxo/Fe(III) LMCT occur at 14,600 cm)1 and 27,200 cm)1 (e = 3800 M)1 cm)1). When the peroxo group bridges the two Fe(III) ions, the p* HOMOs on the peroxo group are split into in-plane p*r and out-of-plane p*p orbitals. The two main LMCT

170

J.-J. Girerd á F. Banse á A.J. Simaan

Fig. 8. Ligands which have allowed the characterization of Fe(III)OOFe(III) complexes,

named according to literature

bands arise from transition between these two orbitals and the Fe d orbitals [46]. 5.3 Resonance Raman Spectroscopy

The Raman data are summarized in Table 12. The mOO frequency occurs between 848 cm)1 and 900 cm)1.

Characterization and Properties of Non-Heme Iron Peroxo Complexes

171

Fig. 9. X-ray diffraction structure of Fe2(l-O2)(l-O2CCH2Ph)2{HB(pz0 )3}2 (from Kim and

Lippard [43])

Fig. 10. X-ray diffraction structure of [Fe2(l-O2)(Ph-bimp)(OBz)]2+ (found by Ookubo et al.

[44])

This frequency is higher than in the Fe(III)OOH and Fe(III)O2 systems (Fig. 4). In Fig. 4 one may also notice that the zone occupied by the FeOOFe systems is ill-de®ned. The FeAO frequency is between 462 cm)1 and 481 cm)1 for the systems which are likely to possess a cis planar conformation [47±49] and is smaller (around 420 cm)1) for the two systems which have a cis-gauche

172

J.-J. Girerd á F. Banse á A.J. Simaan

Fig. 11. X-ray diffraction structure of [Fe2(l-O2)(N-Et-HPTB)(Ph3PO)2]3+ (found by Dong

et al. [45])

Table 11. Wavelength (molar extinction coef®cient) of the peroxo ® Fe(III) LMCT for

Fe(III)OOFe(III) units Complex

kmax (e/M)1 á cm)1)

Refs.

[Fe2(l-O2)(l-O2CCH2Ph)2{HB(pz¢)3}2] [Fe2(l-O2)(l-OBz)2{HB(pz¢)3}2]

694 nm (2650) 682 nm (3450) 679 nm (?), toluene 576 nm (3540), MeCN 588 nm (1500), CH2Cl2 560 (2200), H2O 604 (1600), MeOH 606 (1500), MeOH 600 (1500), MeOH 480 (2370), MeOH 562 nm (3200), CH2Cl2/DMSO 572 nm (2060), CH3CN/DMSO 494 nm (1100), 648 nm (1200) 846 nm (230), CH3CN 500±800 nm (1700), CH3CN 540 nm (2300), CH2Cl2 616 nm (2000), CH2Cl2 618 nm (1000), CH2Cl2 603 nm (1400), CH2Cl2 660 nm (170), THF 670 nm (170), THF

[43] [50]

[44] [54] [55] [55] [55] [56] [56]

540 nm (187), water

[57]

[Fe2(Ph3PO)8(l-O2) á 2H2O]4+ [Fe2(l-O2)(N-Et-HPTB)(OBz)]2+ [Fe2(l-O2)(HPTB)(NO3)2]+ [Fe2(l-O2)(N-Et-HPTB)(NO3)2]+ [Fe2(l-O2)(5,6-Me2-HPTB)(NO3)2]+ [Fe2(l-O2)(5-Me-HXTA)(OBz)]2[Fe2(l-O2)(HPTP)(OBz)]2+ [Fe2(l-O)(l-O2)(6-Me3TPA)2]2+ [Fe2(l-O2)(Ph-bimp)(OBz)]2+ ``[Fe2(Mes2ArCO2)4(l-O2)]'' ``[Fe2(l-O2)(Me4-tpdp)(OBz)]'' ``[Fe2(l-O2)(Me4-tpdp)(CF3COO)]'' ``[Fe2(l-O2)(Me2-tpdp)(CF3COO)]'' ``[Fe2(l-O2)(l-OBz)(l-XDK)(ImH)(OBz)]'' ``[Fe2(l-O2)(l-O2CC(CH3)3)(l-PXDK) (N-MeIm)2(O2CC(CH3)3)]'' ``[Fe2(L1)2(l-O2)]''

[51] [52] [47] [47] [47] [53] [48] [49]

Complex

[Fe2(l-O2)(l-O2CCH2Ph)2{HB(pz¢)3}2] [Fe2(l-O2)(l-OBz)2{HB(pz¢)3}2] [Fe2(Ph3PO)8(l-O2) á 2H2O]4+ [Fe2(l-O2)(HPTB)(NO3)2]+ [Fe2(l-O2)(N-Et-HPTB)(NO3)2]+ [Fe2(l-O2)(5,6-Me2-HPTB)(NO3)2]+ [Fe2(l-O2)(N-Et-HPTB)(OBz)]2+ [Fe2(l-O2)(HPTP)(OBz)]2+ [Fe2(l-O)(l-O2)(6-Me3TPA)2]2+ ``[Fe2(Mes2ArCO2)4(l-O2)]'' ``[Fe2(l-O2)(Me4-tpdp)(OBz)]'' [Fe2(l-O2)(5-Me-HXTA)(OBz)]2± Desaturase RNR mutant Ferritin

16

O16O

16

O16O or 18

18

O18O

O18O (Raman shifts in cm)1)

mOO

mFeO

mFeOas

mOO

mFeO

888 876 882 895 890 888 900 877, 893 848 885 891, 918 884 898 870 851

415 418 445 476 476 474 476 453, 481 462 ± 450 ± 442 458 485

± ± 565 ± ± ± ± ± ± ± 486 ± 490 499 499

842 ()46) 827 ()49) 848 ()34) 854 ()41) 838 ()52) ± 850 ()50) 834 802 ()46) 871 ()14) 857, 889 ± 845 ()53) 824 ()46) 800 ()51)

404 409 ± 457 460 ± 460 444 441 ± 442 ± 425 442 468

Refs.

mFeOas ()11) ()9) ()19) ()16) ()16) ()21) ()8) ()17) ()16) ()17)

± ± ± ± ± ± ± ± ± 479 ± 471 477 487

()7) ()19) ()22) ()12)

[43] [50a] [51] [47] [47] [47] [48] [48] [49] [54] [55] [53] [60] [59] [58]

Characterization and Properties of Non-Heme Iron Peroxo Complexes

Table 12. Raman mOO, mFeO, and mFeOas frequencies in cm)1 for Fe(III)O2Fe(III) with

173

174

J.-J. Girerd á F. Banse á A.J. Simaan

conformation [43, 50a]. In these two last cases, the FeAO frequencies are the lowest known to date for all the systems in this review. In their study, Brunold et al. [46] speci®cally addressed the question of the large value of the OAO stretching and proposed that it is due to mechanical coupling with FeAO stretchings and that intrinsically the strength of the OAO Ê ) is similar to that in oxyhemet (fOO = 3.29 mdyn/A Ê) bond (fOO = 3.1 mdyn/A Ê ) systems. or FeO2 (fOO = 3.02 mdyn/A Raman data have been recently published for enzymes which present a peroxo intermediate which is very likely of the l-1,2-type. In ferritin [58], the following Raman frequencies have been observed (Table 12): mOO = 851 cm)1, mFeOs = 485 cm)1, mFeOas = 499 cm)1. The mOO and mFeOs values are included in Fig. 4. They are close to the cis planar model system prepared with 6-Me3TPA [49]. Data reported for a RNR mutant [59] (mOO = 870 cm)1, mFeOs = 458 cm)1, mFeOas = 499 cm)1) and a Desaturase [60] (mOO = 898 cm)1, mFeOs = 442 cm)1, mFeOas = 490 cm)1) are different from each other. In Fig. 4 they are intermediate between values measured for model complexes, although one may argue that the signals for desaturase are closer to the cis-gauche model systems and that for the RNR mutant is closer to cis planar. It seems that the cis-gauche Fe2(O2)(O2CCH2Ph)2{HB(pz0 )3}2 chemical model is very close to the peroxo intermediate in Desaturase.

6 Conclusion Iron-peroxo or hydroperoxo systems are important intermediates between the initial Fe(II)-dioxygen adduct and the ``activated'' form of the catalytic site in many enzymes [2, 5]. In Scheme 15, mononuclear iron systems in which peroxo forms have been identi®ed to date are summarized. The systems with a question mark have been proposed to imply the intermediate of the type indicated but to our knowledge this intermediate has not been spectroscopically characterized. The g2-peroxo form has been identi®ed with heme systems [61, 62], EDTA4) [39±41], and polydentate aminopyridine ligands [8, 11, 42]. The g1peroxo may exist in some heme systems but it is not yet proven spectroscopically. These peroxo forms can be protonated. In case of aminopyridine or BLM, the monoprotonated form Fe(III)(g1-OOH) can be isolated and spectroscopically studied (see above). In heme systems or EDTA4), the Fe(III)(g1-OOH) seems basic and undergoes a second protonation to lead to Fe(V)O, which prohibits its observation. With aminopyridine and BLM it is possible that a second protonation occurs but information is lacking. The low spin nature of Fe(III)(g1-OOH) seems decisive in explaining the strength of the FeAO bond and the weakness of the OAO bond, preludes to the evolution of Fe(III)(g1-OOH) into a Fe(V)O type structure by breakage of the OAO bond.

Characterization and Properties of Non-Heme Iron Peroxo Complexes

175

Scheme 15.

Although to date spectroscopic evidence has been given only for Fe(III) (g2-OO) and Fe(III)(g1-OOH) species, the existence of the ``complementary'' Fe(III)(g1-OO) and Fe(III)(g2-OOH) species is quite possible and their spectroscopic detection would be very valuable. Finally, the peroxo bridged Fe(III)OOFe(III) systems are also fascinating. How these change to the Fe(IV)O2Fe(IV) form which has been shown to be the active species in MMO is a very interesting question [63]. One possibility is an internal rearrangement which would be analogous to that which is observed in copper chemistry (see the chapter by Tolman in this book). Another possibility has been suggested recently by Brunold et al. (Scheme 16) [46].

Scheme 16.

It implies the protonation of a cis l-1,2 peroxo group to give a l-g1 hydroperoxo system which upon a second protonation leads to the Fe(IV)Fe(IV) dinuclear site. An alternative leading to incorporation into the

176

J.-J. Girerd á F. Banse á A.J. Simaan

Fe(IV)O2Fe(IV) form, of only one of the two oxygen atoms of O2, has recently been proposed by Lee and Lipscomb [64]. Certainly spectroscopy on chemical models and proteins will allow one to track the details of this transformation. Note added in Proof. During the preparation of this manuscript one more example of Fe-Peroxo Unit Fe(III)O2 with the N4Py ligand has been submitted [65]. It has appeared after the completion of the present review.

7 References 1. Lippard SJ, Berg JM (1994) Principles of bioinorganic chemistry. University Science Books, Mill Valley 2. Sono M, Roach MP, Coulter ED, Dawson JH (1996) Chem Rev 96: 2841 3. Ferguson-Miller S, Babcock GT (1996) Chem Rev 96: 2889 4. Wallar BJ, Lipscomb JD (1996) Chem Rev 96: 2625 5. Que L, Ho RYN (1996) Chem Rev 96: 2607 6. Sam JW, Tang XJ, Peisach J (1994) J Am Chem Soc 116: 5250 7. Mizoguchi TJ, Lippard SJ (1998) J Am Chem Soc 120: 11022 8. Simaan AJ et al., Eur J Inorg Chem, in press 9. Bernal I, Jensen IM, Jensen KB, McKenzie CJ, Toftlund H, Tuchagues JP (1995) J Chem Soc Dalton Trans. 3667 10. Mialane P, Nivorojkine A, Pratviel G, AzeÂma L, Slany M, Godde F, Simaan A, Banse F, Kargar-Grisel T, Bouchoux G, Sainton J, Horner O, Guilhem J, Tchertanova L, Meunier B, Girerd JJ (1999) Inorg Chem 38: 1085 11. Jensen KB, McKenzie CJ, Nielsen LP, Pedersen JZ, Svendsen HM (1999) Chem Commun 1313 12. Lubben M, Meetsma A, Wilkinson EC, Feringa B, Que L (1995) Angew Chem Int Ed Engl 34: 1512 13. Ho RYN, Roelfes G, Feringa BL, Que L (1999) J Am Chem Soc 121: 264 14. Roelfes G, Lubben M, Chen K, Ho RYN, Meestma A, Genseberger S, Hermant RM, Hage R, Mandal SK, Young VG, Zang Y, Kooijman H, Spek AL, Que L, Feringa BL (1999) Inorg Chem 38: 1929 15. deVries ME, LaCrois RM, Roelfes G, Kooijman H, Spek AL, Hage R, Feringa BL (1997) Chem Commun 1549 16. Lippai I, Magliozzo RS, Peisach J (1999) J Am Chem Soc 121: 780 17. Sauer-Masarwa A, Herron N, Fendrick CM, Busch DH (1993) Inorg Chem 32: 1086 18. Kim C, Chen K, Kim J, Que L (1997) J Am Chem Soc 119: 5964 19. Burger RM, Horwitz SB, Peisach J, Wittenberg JB (1979) J of Biol Chem 254: 12,299 20. Sobolev AP, Babushkin DE, Talsi EP (1996) Mendeleev Commun 236 21. Grif®th JS (1961) The theory of transition metal ions. Cambridge University Press, London 22. Taylor CPS (1977) Biochimica et Biophysica Acta 491: 137 23. Veselov A, Sun H, Sienkiewicz A, Taylor H, Burger RM, Scholes CP (1995) J Am Chem Soc 117: 7508 24. Hori H (1971) Biochimica et Biophysica Acta 251: 227 25. (a) Brunold TC, Solomon EI (1999) J Am Chem Soc 121: 8277; (b) Brunold TC, Solomon EI (1999) J Am Chem Soc 121: 8288 26. Shiemke AK, Loehr TM, Sanders-Loehr J (1984) J Am Chem Soc 106: 4951 27. Kurtz DM, Shriver DF, Klotz IM (1976) J Am Chem Soc 98: 5033 28. Guajardo RJ, Hudson SE, Brown SJ, Mascharak PK (1993) J Am Chem Soc 115: 7971 29. Chen K, Que L (1999) Chem Commun 1375 30. Chen K, Que L (1999) Angew Chem Int Ed Engl 38: 2227

Characterization and Properties of Non-Heme Iron Peroxo Complexes

177

31. Wada H, Ogo S, Watanabe Y, Mukai M, Kitagawa T, Jitsukawa K, Masuda H, Einaga H (1999) Inorg Chem 38: 3592 32. Zang Y, Elgren TE, Dong Y, Que L (1993) J Am Chem Soc 115: 811 33. Zang Y, Kim J, Dong Y, Wilkinson EC, Appleman EH, Que L (1997) J Am Chem Soc 119: 4197 34. Lange SJ, Miyake H, Que L (1999) J Am Chem Soc 121: 6330 35. Kim J, Larka E, Wilkinson EC, Que L (1995) Ang. Chem Int Ed Engl 34: 2048 36. MeÂnage S, Wilkinson EC, Que L, Fontecave M (1995) Angew Chem Int Ed Engl 34: 203 37. Nguyen C, Guajardo RJ, Mascharak PK (1996) Inorg Chem 35: 6273 38. MacFaul PA, Ingold KU, Wayner DDM, Que L (1997) J Am Chem Soc 119: 10,594 39. Walling C, Kurz M, Schugar HJ (1970) Inorg Chem 9: 931 40. Ahmad S, McCallum JD, Shiemke AK, Appelman EH, Loehr TM, Sanders-Loehr J (1988) Inorg Chem 27: 2230 41. Neese F, Solomon EI (1998) J Am Chem Soc 120: 12,829 42. Simaan AJ, Banse F, Mialane P, Boussac A, Un S, Kargar-Grisel T, Bouchoux G, Girerd JJ (1999) Eur J Inorg Chem 993 43. Kim K, Lippard SJ (1996) J Am Chem Soc 118: 4914 44. Ookubo T, Sugimoto H, Nagayama T, Masuda H, Sato T, Tanaka K, Maeda Y, Okawa H, Hayashi Y, Uehara A, Suzuki M (1996) J Am Chem Soc 118: 701 45. Dong Y, Yan S, Young VG, Que L (1996) Angew Chem Int Ed Engl 35: 618 46. Brunold TC, Tamura N, Kitajima N, Moro-oka Y, Solomon EI (1998) J Am Chem Soc 120: 5674 47. Brennan BA, Chen Q, Juarez-Garcia C, True AE, O0 Connor CJ, Que L (1991) Inorg Chem 30: 1937 48. Dong Y, MeÂnage S, Brennan BA, Elgren TE, Jang HG, Pearce LL, Que L (1993) J Am Chem Soc 115: 1851 49. Dong Y, Zang Y, Shu L, Wilkinson EC, Que L, Kauffmann K, MuÈnck E (1997) J Am Chem Soc 119: 12,683 50. (a) Kitajima N, Fukui H, Moro-oka Y (1990) J Am Chem Soc 112: 6402; (b) Kitajima N, Tamura N, Amagai H, Fukui H, Moro-Oka Y, Mizutani Y, Kitagawa T, Mathur R, Heerwegh K, Reed CA, Randall CR, Que L, Tatsumi K (1994) J Am Chem Soc 116: 9071 51. Sawyer DT, McDowell MS, Spencer L, Tsang PKS (1989) Inorg Chem 28: 1166 52. MeÂnage S, Brennan BA, Juarez-Garcia C, MuÈnck E, Que L (1990) J Am Chem Soc 112: 6423 53. Murch BP, Bradley FC, Que L (1986) J Am Chem Soc 108: 5027 54. Hagadorn JR, Que L, Tolman WB (1998) J Am Chem Soc 120: 13,531 55. Hayashi Y, Kayatani T, Sugimoto H, Suzuki M, Inomata K, Uehara A, Mizutani Y, Kitagawa T, Maeda Y (1995) J Am Chem Soc 117: 11,220 56. Herold S, Lippard SJ (1997) J Am Chem Soc 119: 145 57. Kimura E, Kodama M, Machida R, Ishizu K (1982) Inorg Chem 21: 595 58. MoeÈnne-Loccoz P, Krebs C, Herlihy K, Edmonson, DE, Theil EC, Huynh BH, Loehr TM (1999) Biochemistry 38: 5290 59. MoeÈnne-Loccoz P, Baldwin J, Ley BA, Loehr TM, Bollinger JM (1998) Biochemistry 37: 14,659 60. Broadwater JA, Ai J, Loehr TM, Sanders-Loehr J, Fox BG (1998) Biochemistry 37: 14,664 61. Friant P, Goulon J, Fischer J, Ricard L, Schappacher M, Weiss R, Momenteau M (1985) Nouveau J Chimie 9: 33 62. Burstyn JN, Roe JA, Miksztal AR, Shaevitz BA, Lang G, Valentine JS (1988) J Am Chem Soc 110: 1382 63. Shu L, Nesheim JC, Kauffmann K, MuÈnck E, Lipscomb JD, Que L (1997) Science 275: 515 64. Lee SY, Lipscomb JD (1999) Biochemistry 38: 4423 65. Ho RYN, Roelfes G, Hermant R, Hage R, Feringa BL, Que L (1999) Chem Commun 2161

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases Allan G. Blackman1, William B. Tolman2 1

2 1 2

Department of Chemistry, University of Otago, P.O. Box 56, Dunedin, New Zealand Department of Chemistry and Center for Metals in Biocatalysis, University of Minnesota, 207 Pleasant St., SE, Minneapolis, MN 55455, USA E-mail: [email protected] E-mail: [email protected]

Copper proteins mediate both the transport and activation of dioxygen in a number of biological systems. The active sites of these proteins comprise mononuclear, dinuclear, and trinuclear copper centers, with the copper ions displaying a variety of coordination numbers and stereochemistries. Information regarding the mechanism by which these proteins activate dioxygen has been obtained by studies of the reactions of small molecule model copper complexes with dioxygen and its derivatives. Superoxo, peroxo, and bis(l-oxo) intermediates in these reactions have recently been characterized by X-ray crystallography and this article concentrates on the structures of these intermediates, along with several Cu/ O2 complexes that have been well characterized by spectroscopic methods. The oxygenasetype reactivities of a number of copper complexes on reaction with dioxygen are also discussed. Keywords: Copper complexes, Dioxygen, X-ray structures, C-H activation, Bioinorganic

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

2

Structures of Copper-Dioxygen and Copper-Oxo Model Complexes . . . . . . . . . . . . . . . . . . . . . 182

2.1 2.1.1 2.1.2 2.2 2.2.1 2.2.2 2.2.3 2.2.4 2.3 2.4

Mononuclear Complexes . . . . . . . . . . . . . . . . . Superoxide Complexes . . . . . . . . . . . . . . . . . . Alkylperoxide, Acylperoxide, and Hydroperoxide Complexes . . . . . . . . . . . . Binuclear Complexes . . . . . . . . . . . . . . . . . . . . l-1,1-Acylperoxo and Hydroperoxo Complexes . l-g1:g1-Peroxo Complexes . . . . . . . . . . . . . . . . l-g2:g2-Peroxo Complexes . . . . . . . . . . . . . . . . Bis(l-Oxo) Complexes . . . . . . . . . . . . . . . . . . . Trinuclear Complexes . . . . . . . . . . . . . . . . . . . Tetranuclear Complexes . . . . . . . . . . . . . . . . .

3

Reactivity of Oxygenase Model Complexes . . . . . . . . . . . . . . . . 201

3.1 3.2

Aromatic Oxygenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202 Aliphatic Oxygenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203

. . . . . . . . . . . . 182 . . . . . . . . . . . . 183 . . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

. . . . . . . .

188 190 190 191 193 195 199 200

Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

180

A.G. Blackman á W.B. Tolman

4

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

5

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208

List of Abbreviations BArF DMF EPR ESI EXAFS LMCT MCD Me2im pzH SQUID UV/vis

tetrakis[3,5-bis(tri¯uoromethyl)phenyl]borate dimethylformamide electron paramagnetic resonance electrospray ionization extended X-ray absorption ®ne structure ligand-to-metal charge transfer magnetic circular dichroism 1,2-dimethylimidazole pyrazole superconducting quantum interference device ultra-violet/visible

1 Introduction The biological importance of copper has been realized since the early years of the nineteenth century, when copper was ®rst isolated from a number of plant species, although its exact function was not understood at that time [1]. Since then, copper proteins and enzymes have been identi®ed in a variety of bacteria, fungi, plant and animal species, mediating such processes as electron transfer, dioxygen transport, dioxygen activation, reduction of nitrogen oxides to elemental nitrogen, and the disproportionation of superoxide ion. Of particular importance are hemocyanin, a dioxygen-carrier protein, and the copper oxidases and oxygenases, which either couple substrate oxidation to the reduction of dioxygen to hydrogen peroxide or water (Eq. 1) or incorporate O2-derived oxygen atoms into substrate (Eq. 2), respectively:

…1†

…2†

Detailed structural information on these proteins has been obtained through X-ray crystallographic and spectroscopic studies, revealing similarities in the types of ligands bound to the active site copper ions (e.g., preponderance of histidine imidazolyls) but great diversity in the metal site nuclearities and

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

181

Fig. 1A±C. X-ray structures of the active sites of: A galactose oxidase at pH 4.5 (PDB

Identi®cation Code 1GOF) [2]; B oxyhemocyanin (PDB Identi®cation Code 1NOL) [3]; C ascorbate oxidase (PDB Identi®cation Code 1AOZ) [4]

coordination geometries (cf. selected examples in Fig. 1) [2±4]. These structural differences are paralleled by differences in mechanism and function. Thus, while the presence of one or more reduced, histidine-coordinated Cu(I) sites that react with atmospheric dioxygen is a common feature, the subsequent course of dioxygen reduction and substrate oxidation varies considerably among hemocyanin and the copper oxidases and oxygenases [5]. With the goal of unraveling important structure/function relationships pertinent to the larger problem of dioxygen binding and activation by metallobiomolecules [6±9], synergistic biochemical and synthetic modeling approaches have been taken in order to understand how dioxygen interacts with the reduced sites of these proteins and what factors control the fate of the resultant intermediates. Dioxygen may be envisaged to bind to a single metal ion following 1- or 2-electron reduction as either superoxide (O2 ) or peroxide (O22 ), and can coordinate in an ``end-on'' g1 or ``side-on'' g2 fashion (Fig. 2). Protonation may occur to yield a hydroperoxide ligand, for example, or subsequent coordination of the bound ligand to a second copper ion can result in a number of bridging modes, such as l-g1, l-g1:g1 (cis or trans), or l-g2:g2. Still further bonding modes are potentially available in tri- and tetranuclear complexes. The OAO bond may also be cleaved on coordination to the copper ion, resulting in complexes containing l-oxo bridges derived from dioxygen.

182

A.G. Blackman á W.B. Tolman

Fig. 2. Possible coordination modes of dioxygen derivatives in mono- and polynuclear

copper complexes

Interconversions among the diverse structures represent possible steps in the mechanisms traversed by the oxidase and oxygenase enzymes. In this review we concentrate on recent efforts to obtain fundamental chemical information on the structures, properties, and reactivities of the possible copper-dioxygen adducts and derived species via studies of the reactivity of copper complexes with O2, H2O2, or related species. While such studies are often complicated by the fact that copper complexes of superoxide and peroxide are generally relatively unstable, the combined use of lowtemperature handling techniques, ligand design principles, and improved analytical methods have recently led to great advances. Most importantly, Xray crystal structures of reactive model complexes have facilitated detailed correlation of structural and spectroscopic features, ultimately enabling entry into biorelevant mechanistic studies. Here we will emphasize complexes that have been structurally well de®ned by X-ray diffraction methods. These are listed in Table 1 along with important metric and spectroscopic data, while Fig. 3 gives the structures of the ligands discussed in this article. Recent reviews on systems characterized by other methods or on related topics are also available [10±19].

2 Structures of Copper-Dioxygen and Copper-Oxo Model Complexes 2.1 Mononuclear Complexes

The initial interaction of a dioxygen molecule with a copper ion in a metalloenzyme undoubtedly results in monodentate g1 or ``side-on'' g2 coordination of a reduced dioxygen species. Thus, the synthesis and characterization of mononuclear copper complexes containing superoxide or peroxide and their derivatives is important in order to gain insight into the initial O2-binding process. Such syntheses have proved to be non-trivial due to

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

183

a propensity for either attack of a second Cu(I) ion to yield dimeric species or for rapid decomposition to occur, so few mononuclear Cu/O2 complexes have been unequivocally structurally characterized. In a strategy with ample precedent in the porphyrin literature [20], dimerization reactions may be inhibited by using sterically bulky multidentate ligands. In some cases, these ligands may also provide a protected cleft in which the bound O2-species may be stabilized by interactions with ligand substituents. This approach has been increasingly exploited over the past decade, resulting in several reports of the successful synthesis and characterization of important mononuclear Cu/O2 complexes. 2.1.1 Superoxide Complexes Monodentate copper-superoxide complexes [LCuO2]+ (L = TMPA or BQPA) have been characterized spectroscopically and shown by kinetic studies to be precursors to dimeric peroxo species, but isolation and structural characterization has not been possible [21, 22]. The spectroscopic characterization of the deep-red diamagnetic superoxo complex [Cu(L1H2)(O2)(NEt3)] obtained on treatment of a Cu(I) precursor with O2 at )50 °C to )70 °C has also recently been reported [23]. In an exciting paper, reaction of a TMPA derivative, tppa, with [Cu(CH3CN)4]ClO4 and O2 in MeOH at )80 °C was reported to yield the monodentate superoxide complex [Cu(tppa)(O2)]ClO4 [24]. The electronic spectrum of the reaction mixture consisted of two bands in the d-d region at 657 nm and 803 nm and a shoulder at 315 nm which was assigned to a O2 ® Cu2+ charge transfer transition; these features are similar to those reported previously for the TMPA and BQPA systems [21, 22]. The reaction mixture was EPR silent at )80 °C and this was proposed to be due to strong antiferromagnetic coupling between the unpaired electrons on the Cu atom and the superoxide ligand. No vibrational spectroscopic data to support the superoxide formulation were reported. Storage of the reaction mixture at low temperature for several days gave a crystalline product that was subjected to room-temperature crystal structural analysis. The data were interpreted to show the presence of a monodentate coordinated superoxide ligand stabilized by hydrogen bonding to the amide protons of the tripodal ligand (Fig. 4). However, the structure of the supposed superoxo complex was found by other workers to be incorrect; redetermination using the original X-ray data showed it to be [Cu(tppa)OH]ClO4 [25] (Fig. 5). The correct direct-methods solution showed the position of all non-hydrogen atoms with the exception of the unbound oxygen atom (originally reported to have unreasonably large thermal parameters) and no peak corresponding to this atom was observed in the difference Fourier map after several cycles of least-squares re®nements. Instead, a peak corresponding to the hydrogen atom of a hydroxide ligand was found and the complex was therefore formulated as a unique example of a Cu2+-OH species in which the hydroxide ligand engages in intramolecular hydrogen bonding entirely different from that postulated for the putative superoxide complex.

184 Table 1. Metric and spectroscopic parameters for Cu-dioxygen, oxo and alkyl/acyl/hydroperoxo complexes that have been characterized by X-ray

crystallography

Ê d(CuAO)/A

Ê d(OAO)/A

[Cu(HB(3-t-Bu-5-i-Prpz)3)(O2)]

1.84 (1)

1.22 (3)

[Cu(HB(3,5-i-Pr2pz)3)(OOCMe2Ph)]

1.816 (4) 1.814 (4)c

1.460 (6) 1.454 (6)

[Cu(bppa)(OOH)]ClO4

1.888 (4)

1.460 (6)

[Cu2(XYL-O-)(m-ClC6H4C(O)O2)] (ClO4)2 á CH3CN [{Cu(TPA)}2(O2)](PF6)2 á 5Et2O

1.948 (8) 1.971 (9) 1.852 (5)

1.463 (12)

3.197

1.432 (6)

4.359 (1)

[{Cu(HB(3,5-i-Pr2pz)3)}2(O2)]

1.903 1.927 1.808 1.803 1.823 1.836 1.824 1.827

1.412 (12)

3.560 (3)

2.287 (2)

2.794 (2)

2.351 (2)

2.783 (1)

[{Cu(d21-Bn3TACN)}2(O)2] (SbF6)2 á 7(CH3)2CO á 2CH3CN [Cu2(d28-i-Pr4DTNE)(l-O)2] (SbF6)2 á 3CH2Cl2

(11) (9) (5) (5) (4) (4) (4) (4)

Ê d(Cu    Cu)/A

kmax/nm (e/M)1 cm)1)

m(OAO)/cm)1b

Refs.

352 510 348 569 811 380 660 830 395 650 435 524 615 1035 349 551 318 430 316 414

1111 (1062)

[28]

Not reported

[32]

856 (810)

[36]

Not reported

[38]

832 (788)

[44, 46]

741 (698)

[31, 52]

602,608 (583) (CuAO) 600 (582,574) (CuAO)

[64]

(2330) (230) (4000) (3600) (400) (890) (150) (250) (5500±5900) (420) (1700) (11300) (5800) (180) (21000) (790) (12000) (14000) (13000) (14000)

[65]

A.G. Blackman á W.B. Tolman

Complexa

a b c d

2.344 (1)

2.743 (1)

306 (21000) 401 (28000)

610 (587) (CuAO)

[67]

2.37

2.641 (3) 2.704 (3)

290 355 480 620

Not reported

[85]

1.453 (4)

2.994 3.030 4.184 4.317

284 (15900) 384 (9700) 587 (610)

878 (841)

[87, 90]

(2) (2) (3) (2)

Ligand abbreviations are given in Fig. 3. Numbers in parentheses refer to the 18O-labeled derivative. Two independent molecules in the asymmetric unit. Two independent molecules in the asymmetric unit; data only reported for one.

(12500) (15000) (1400) (800)

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

[{Cu(LME)}2(O)2](CF3SO3)2 á 4CH2Cl2 1.814 (6) 1.809 (6) 1.796 (6) 1.804 (6) [(LTM)3Cu3O2](CF3SO3)3d 1.83 (1) 1.83 (1) 1.98 (1) 2.01 (1) 2.01 (1) 2.04 (2) [Cu4(L2)2(O2)(OMe)2(ClO4)] 1.961 (2) ClO4 á MeOH 1.939 (2)

185

186

A.G. Blackman á W.B. Tolman

Fig. 3. Ligand structures and abbreviations

As part of this same study, the hydroxide complex was synthesized independently and the X-ray structure of this material was found to be identical to that obtained from the previously published data, con®rming unequivocally that the original formulation of the crystalline material as a superoxide complex was in error. Debate about the nature of the solution species in the reaction mixture still continues [26, 27], but in the absence of corroborating vibrational spectral data or a bona ®de X-ray crystal structure,

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

187

Fig. 4. X-ray structure of the purported monodentate superoxide complex cation

Ê [Cu(tppa)(O2)]+ [24]. Selected bond lengths are given in A

Ê Fig. 5. X-ray structure of the [Cu(tppa)OH]+ cation [25]. Selected bond lengths are given in A

the original report of the generation of [Cu(tppa)(O2)]ClO4 must remain under suspicion. A crystallographically characterized example of superoxide coordinated to Cu in an g2 fashion was reported in 1994 [28]. The complex was prepared by exposure of a solution of [Cu(HB(3-t-Bu-5-i-Prpz)3)(DMF)] to O2 in CH2Cl2 at )50 °C. Manometry suggested a 1:1 Cu:O2 ratio, and reversible O2 binding was demonstrated in toluene by vacuum cycling. A red-brown solid was obtained from the reaction mixture at )20 °C, elemental analysis of which suggested the formula [Cu(HB(3-t-Bu-5-i-Prpz)3)(O2)]. The IR spectrum of the solid exhibited a band at 1112 cm)1, which was assigned to m(OAO) of a superoxide ion. Also, the resonance Raman spectrum of the complex prepared using 16 O2 showed a band at 1111 cm)1 which shifted the expected amount for an OAO oscillator to 1062 cm)1 on isotopic substitution with 18 O2 . The solid was diamagnetic, as evidenced by solid-state SQUID magnetic susceptibility measurements and by its sharp 1H NMR spectrum at )40 °C, suggesting a

188

A.G. Blackman á W.B. Tolman

strong antiferromagnetic interaction between the superoxide ligand and the copper atom. A poor quality data set was obtained from weakly diffracting crystals which nonetheless con®rmed the identity of the complex (Fig. 6). A distorted square-pyramidal geometry applies about the copper atom, with the ``side-on'' bound superoxide ligand occupying two basal plane positions. While g1-binding appears to be preferred when the supporting ligand is tetradentate (see above), g2 binding is favored by the tridentate supporting ligand in [Cu(HB(3-t-Bu-5-i-Prpz)3)(O2)]. The tendency for CuII to adopt 5coordinate geometries thus is a key structural determinant in Cu/O2 chemistry. Ê con®rms its formulation as a superoxide The OAO bond distance of 1.22(3) A complex, being very similar to that found in the closely related superoxo Ê ) [29] and much shorter complex [Co(HB(3-t-Bu-5-Mepz)3)(O2)] (1.262(8) A than the OAO bond distance in the side-on bound peroxo complex [Mn(3,5-iÊ ) [30]. The successful isolation of Pr2pzH)(HB(3,5-i-Pr2pz)3)(O2)] (1.43(1) A monomeric [Cu(HB(3-t-Bu-5-i-Prpz)3)(O2)] was presumably dependent on the bulky nature of the supporting N-donor ligand which obviated formation of a superoxo- or peroxo-bridged dimer. The presence of the tert-butyl group in the 3-position appears essential to stop dimerization, as the same reaction using either HB(3,5-i-Pr2pz)3 or HB(3,5-Ph2pz)3 as the supporting ligand gives l-g2:g2-peroxo species (see below) [31]. 2.1.2 Alkylperoxide, Acylperoxide, and Hydroperoxide Complexes The syntheses of the monodentate alkylperoxide complexes [Cu(HB(3,5-iPr2pz)3)(OOR)] (R = t-Bu or CMe2Ph) were reported in 1993 [32]. The complexes were obtained from reaction of the bis(lm-hydroxo)-bridged dimer [{Cu(HB(3,5-i-Pr2pz)3)}2(OH)2] with excess tert-butyl- or cumyl-hydroperoxide in pentane at )20 °C to give dark blue microcrystalline solids that were stable below )20 °C for a week, but which decomposed at room temperature. The X-ray structure of the compound with R = CMe2Ph (Fig. 7) con®rmed the

Fig. 6. X-ray structure of the g2-superoxo complex [Cu(HB(3-t-Bu-5-i-Prpz)3)(O2)] [28].

Ê Selected bond lengths are given in A

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

189

Fig. 7. X-ray structure of the alkylperoxide complex [Cu(HB(3,5-i-Pr2pz)3)(OOCMe2Ph)]

Ê [32]. Selected bond lengths are given in A

monomeric formulation of the complex and showed a distorted tetrahedral Ê ) CuAO bond and an geometry about the copper ion, with a short (1.816(4) A Ê . The OAO bond distance in the cumylperoxide ligand of 1.460(6) A monodentate cumylperoxide ligand is coordinated in a bent fashion, with a CuAOAO bond angle of 112.1(4)°. These structural parameters are very similar to those observed in a series of crystallographically characterized Co(III) alkylperoxo complexes [33±35]. The only structurally characterized copper-hydroperoxo complex was reported in 1998 using the sterically hindered tripodal ligand bppa (Fig. 8) [36]. Addition of a large (20-fold) excess of aqueous H2O2 to acetonitrile solutions of [Cu(bppa)]ClO4 or [Cu(bppa)(CH3CO2)]ClO4 gave [Cu(bppa) (OOH)]ClO4 which was characterized by X-ray crystallography, EPR, electronic and resonance Raman spectroscopy, and ESI mass spectrometry. The copper atom adopts a trigonal bipyramidal geometry, and the hydroperoxide ligand is bound in a bent fashion with a CuAOAO bond angle of 114.5°. The Ê ) is only slightly shorter than that found in H2O2 OAO bond length (1.460(6) A

Fig. 8. X-ray structure of the hydroperoxide complex cation [Cu(bppa)(OOH)]+ [36].

Ê Selected bond lengths are given in A

190

A.G. Blackman á W.B. Tolman

Ê ), supporting the assignment of this as a hydroperoxide rather than a (1.490 A superoxide ligand. The remarkable stability of [Cu(bppa)(OOH)]ClO4 (no change in the electronic absorption spectrum over one month at room temperature in MeCN) appears to be due to a combination of the bulkiness of the hydrophobic bppa ligand and the presence of two internal hydrogen bonds from the amide N-H protons to the coordinated oxygen atom of the hydroperoxide ligand. The electronic spectrum of [Cu(bppa)(OOH)]ClO4 consists of two low-intensity d-d bands at 830 nm and 660 nm and a more intense band around 380 nm which was assigned to a ligand-to-metal chargetransfer transition. Of critical importance for con®rming the structure of the complex was the resonance Raman spectrum (excitation wavelength 441.6 nm), which showed a peak at 856 cm)1 that shifted to 810 cm)1 when H2 18 O2 was used in the synthesis. The EPR spectrum of [Cu(bppa)(OOH)]ClO4 is typical of trigonal bipyramidal Cu2+ and the ESI mass spectrum shows peaks at m/z 584 ([Cu(bppa)(OOH)]+) and 784 ({[Cu(bppa)(OOH)](ClO4)2})) which moved to m/z 588 and 788 respectively when H2 18 O2 was used. Synthesis of the monodentate mononuclear acylperoxo complexes and [Cu(Me2im)3(m[Cu(TMPA)(m-ClC6H4C(O)O2)](PF6) á 0.5CH2Cl2 ClC6H4C(O)O2)](PF6) was achieved by treatment of the corresponding peroxo dimers with m-CPBA at )90 °C [37]. Magnetic and EPR data obtained on the room-temperature stable complexes were consistent with their formulation as monomers, while IR data suggested monodentate, rather than bidentate coordination of the peracid ligand. 2.2 Binuclear Complexes

By far the most common copper-peroxo and -oxo complexes to be structurally characterized are binuclear, the reduction of dioxygen by two copper ions to yield (peroxo)dicopper(II) or bis(oxo)dicopper(III) species being particularly thermodynamically favorable. As in the case of the mononuclear compounds described above (Sect. 2.1), the number, disposition, and type of N-donors in the supporting ligand as well as the steric in¯uences of the ligand substituents have been shown to be important for controlling the course of the CuI/O2 reactions and allowing isolation of discrete, crystalline complexes. 2.2.1 l-1,1-Acylperoxo and Hydroperoxo Complexes Although not a species derived directly from dioxygen, the synthesis and characterization in 1987 of the acylperoxo dicopper complex [Cu2(XYL-O-)(m-ClC6H4C(O)O2)](ClO4)2 á CH3CN (Fig. 9) represented a major advance in the ®eld of copper-dioxygen chemistry as this was the ®rst reported X-ray structure of a copper-peroxo species of any type [38]. The complex was prepared at )80 °C either by treatment of the hydroxo-bridged dimer [Cu2(XYL-O)(OH)]2+ with m-CPBA or acylation of the peroxo-bridged dimer [Cu2(XYL-O)(O2)]+ (see below) with m-chlorobenzoyl chloride. The

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

191

Fig. 9. X-ray structure of the acylperoxide complex cation [Cu2(XYL-O-)(m-ClC6H4C(O)

Ê O2)]2+ [38]. Selected bond lengths are given in A

crystal structure of the complex showed the acylperoxide ligand to be bound such that the terminal oxygen atom bridged the two copper ions, resulting in a Ê. Cu  Cu separation of 3.197 A Although no X-ray structures are available, a strong case for the existence of l-1,1 hydroperoxo dimers has been made on the basis of spectroscopic data. Treatment of the hydroxo-bridged dimer [Cu2(XYL-O-)(OH)]2) with excess H2O2 at )80 °C, reaction of the peroxo-bridged dimer [Cu2(XYL-O-)(O2)]+ with HBF4 á Et2O at )80 °C and oxygenation of [Cu2(XYL-OH)]2+ at )80 °C all yielded the same product, identi®ed as the l-1,1-hydroperoxo species [Cu2(XYL-O-)(OOH)]2+ on the basis of the similarity of its spectral features with those of the structurally characterized acylperoxo complex [Cu2(XYL-O-) (m-ClC6H4C(O)O2)](ClO4)2 á CH3CN [39]. Both the acylperoxo and hydroperoxo complexes exhibit intense absorptions at 395 nm in the UV/vis spectrum, as well as weaker d-d bands at longer wavelengths. Further support for the l-1,1-hydroperoxo formulation of [Cu2(XYL-O-)(OOH)]2+ was obtained from comparison of EXAFS data for this complex with those for the structurally characterized complex [Cu2(XYL-O-)(OH)]2+ [40]. While this hydroperoxo complex was only stable at low temperatures, slight modi®cation of the XYL-O ligand [41] allowed isolation of a room-temperature stable hydroperoxo complex [Cu2(UN-O-)(OOH)](PF6)2 á 1/2Et2O [42]. A band at 892 cm)1 in the resonance Raman spectrum of this complex was assigned to m(OAO) and was observed to shift to 840 cm)1 on substitution with 18O2 [43]. 2.2.2 l-g1:g1-Peroxo Complexes The ®rst copper-dioxygen complex to be structurally characterized was reported in 1988 [44]. Reaction of [Cu(TMPA)(RCN)]PF6 (R = Me, Et) with O2 at )80 °C in EtCN or CH2Cl2 gave a color change from orange to deep purple.

192

A.G. Blackman á W.B. Tolman

Manometry showed the reaction stoichiometry to be Cu:O2 = 2:1 and the binding of O2 was shown to be reversible by vacuum/warming cycling experiments. Crystals of [{Cu(TMPA)}2(O2)](PF6)2 á 5Et2O were obtained on storage of the reaction mixture at )85 °C (Fig. 10). The structure consists of two Cu(TMPA)2+ moieties bridged by peroxide in Ê . This binding mode a trans-1,2 fashion with a CuACu distance of 4.359 A apparently is preferred because of the tetradentate supporting N-donor ligand; due to the aforementioned propensity for Cu(II) to adopt 5-coordinate geometries (here, distorted trigonal bipyramidal), monodentate coordination Ê of peroxide is favored. Importantly, the OAO bond length of 1.432(6) A con®rms the assignment of a bridging peroxide ligand. The EPR-silence of the product is indicative of a strong antiferromagnetic coupling between the copper atoms mediated by the bridging peroxo ligand. Full spectroscopic characterization for the complex was reported in subsequent papers [45, 46]. Three major bands were observed in the UV/vis spectrum at 615 (e = 5800 M)1 cm)1), 524 (e = 11300 M)1 cm)1) and 435 nm (e = 1700 M)1 cm)1), and these were assigned to O22 ® Cu2+ charge transfer transitions. A less intense band at 1035 nm (e = 180 M)1 cm)1) was assigned to a d-d transition. Bands in the resonance Raman spectrum at 832 cm)1 and 561 cm)1 were observed to shift to 788 cm)1 and 535 cm)1 on substitution of 16 O2 by 18O2. These bands were assigned to m(OAO) and a symmetric m(CuAO), respectively, on the basis of their isotope shifts and characteristic frequencies. The synthesis of [{Cu(TMPA)}2(O2)](PF6)2 represented an important advance in copper-dioxygen chemistry as it showed for the ®rst time that the correct choice of ligand combined with low-temperature conditions could allow the isolation of copper-dioxygen species. Moreover, the de®nitive correlation of spectroscopic and X-ray crystal structural data for this compound has allowed subsequent identi®cation of the trans-1,2-O22 unit in other copper complexes, all of which have in common tetradentate ligation of an N-donor supporting ligand [47±49]. Of great interest amongst these is the remarkably stable purple species formed on oxygenation of the dicopper(I) complex of the macrocyclic ligand MEPY22PZ [49]. This complex shows intense bands at 525 nm and 625 nm and a weak shoulder at 438 nm in the

Fig. 10. X-ray structure of the peroxo complex cation [{Cu(TMPA)}2(O2)]2+ [44]. Selected

Ê bond lengths are given in A

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

193

UV/vis spectrum, while the resonance Raman spectrum shows a band assigned to m(OAO) at 844 cm)1 which shifts to 798 cm)1 on substitution with 18O2. The authors proposed this to be an intramolecular trans-1,2-peroxo species, although molecular mechanics calculations showed that formation of an intermolecular species was also feasible. The stability of the peroxo species (t1/2 = 250 s in MeOH at 21 °C) was ascribed to the relative inaccessibility of the macrocyclic cavity in which the metal ions are coordinated. 2.2.3 l-g2:g2-Peroxo Complexes In 1988, the synthesis of a peroxide-bridged Cu2+ complex from the reaction of [{Cu(HB(3,5-Me2pz)3)}2O] with an equimolar amount of H2O2 at )35 °C was reported [50]. The product, proposed to be [{Cu(HB(3,5-Me2pz)3)}2(O2)] on the basis of elemental analysis and mass spectral data, was diamagnetic, giving no EPR signal and exhibiting a sharp 1H NMR spectrum. The resonance Raman spectrum showed a band at 725 cm)1 that shifted to 686 cm)1 when H18 2 O2 was used to prepare the complex. While the isotope shift supported the presence of an OAO bond in the complex, the band energies were rather low for a peroxo unit (typically m(OAO) > 800 cm)1 in transition metal peroxide compounds) [51]. The UV/vis spectrum of the dark purple solid was remarkably similar to that of oxyhemocyanin, with bands at 530 nm (e = 840 M)1 cm)1) and 338 nm (e = 20,800 M)1 cm)1). In view of this similarity and analogous resonance Raman and magnetic properties, the complex was advanced as a structural model of this protein, albeit in the absence of crystal structural data needed to con®rm the mode of binding of peroxide to the metal centers. These critical data were available the following year, when the same group reported the synthesis of [{Cu(HB(3,5-i-Pr2pz)3)}2 (O2)] by both O2 addition to the Cu(I) precursor [Cu(HB(3,5-i-Pr2pz)3)] and reaction of H2O2 with the dimeric Cu(II) complex [{Cu(HB(3,5-i-Pr2pz)3)}2 (OH)2] [52]. The spectral features of the peroxo product were similar to those reported for the dimethyl analogue above, but enhanced stability and crystallinity conferred by isopropyl substitution on the ligands allowed further characterization by X-ray crystallography (Fig. 11).

Fig. 11. X-ray structure of the peroxo complex [{Cu(HB(3,5-i-Pr2pz)3)}2(O2)] [52]. Selected

Ê bond lengths are given in A

194

A.G. Blackman á W.B. Tolman

Ê apart) The structure revealed peroxide bridging the two copper ions (3.56 A 2 2 in a symmetrical, planar g :g fashion with all CuAO distances the same within Ê . Although seen experimental error and an OAO distance of 1.412(12) A previously in complexes of lanthanum [53] and uranium [54], this planar (l-g2:g2-peroxo)dimetal unit was a ®rst in transition metal chemistry. Full spectroscopic characterization of the [{Cu(HB(3,5-R2pz)3)}2(O2)] (R = Me, i-Pr, Ph) complexes has been described, revealing, among other things, close agreement with the properties of oxyhemocyanin [31, 55]. Final con®rmation of the structural congruity between the complexes and the protein active site came with the report of the X-ray structure of oxyhemocyanin [3]. The correct prediction through exploratory synthetic chemistry of the presence of the (l-g2:g2-peroxo)dicopper core in oxyhemocyanin ranks as one of the greatest successes of bioinorganic model chemistry. Since the initial discoveries using the HB(3,5-R2pz)3 ligand system, additional (l-g2:g2-peroxo)dicopper complexes have been identi®ed by spectroscopic means [56±58]. There are also a number of peroxo complexes which, while displaying resonance Raman spectral properties consistent with a l-g2:g2-peroxo formulation (e.g. m(OAO) = 747 cm)1, shifted to 707 cm)1 on substitution by 18O2 [59]) also show slightly anomalous UV/vis spectra. In addition to bands around 360 nm and 530 nm typical of l-g2:g2-peroxo complexes, they also display a moderately intense band in the range 420± 490 nm [60, 61]. This has been proposed to be due to a bent ``butter¯y'' l-g2:g2-peroxo core (Fig. 12), with the extra band being due to an LMCT transition which is formally forbidden in the planar l-g2:g2-peroxo core [62]. The possibility that the extra band was due to the isomeric bis(l-oxo) dicopper(III) core (see below) was discounted on the basis of a resonance Raman excitation pro®le [59]. Interconversion between two different types of peroxo cores has been observed in complexes of the binucleating ligand N4. Stopped-¯ow spectrophotometric studies of the oxygenation of the Cu(I) complex [(N4)Cu2]2+ at )94 °C gave evidence for the initial formation of a complex containing the l-g1:g1-peroxo core, which then rapidly rearranged to the bent ``butter¯y'' l-g2:g2-peroxo core [63]. Kinetic data suggested that this rearrangement could not occur directly, and a mechanism involving the intermediacy of either the Cu(I) starting material or a transient superoxo Cu(II) species was proposed. All of the complexes containing the l-g2:g2-peroxo core incorporate a tridentate N-donor ligand that allows g2-coordination of the peroxide to give 5-coordinate Cu(II) ions, and each has UV/vis spectral features and an

Fig. 12. The bent ``butter¯y'' l-g2:g2-peroxo core

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

195

anomolously low OAO stretching frequency now recognized to be signatures of the (l-g2:g2-peroxo)dicopper unit. In addition to being present in oxyhemocyanin, the weakened OAO bond implied by the generally low m(OAO) value suggests that this bonding motif may lie on the reaction pathway of dioxygen OAO bond cleavage in enzymatic and other catalytic oxidations (see below). 2.2.4 Bis(l-Oxo) Complexes While the presence of Cu(I) and Cu(II) in metalloproteins is well established, it has long been thought that complexes of Cu(III) are unlikely to exist in biological systems due to the high redox potential required to access this oxidation state. However, this view must now be called into question with the recent synthesis and structural characterization of several bis(l-oxo)dicopper(III) complexes [14, 18, 64±67]. In all cases, the complexes were prepared via oxidation of a Cu(I) precursor with molecular oxygen, thus demonstrating the viability of generating high-valent copper species in biological systems. The X-ray structures of three compounds in this class are shown in Figs. 13±15. All three are thermally unstable molecules that decompose via liganddegradation pathways involving rate-controlling CAH bond scission characterized by large H/D kinetic isotope effects (see below). As a result, in order to slow decay suf®ciently so that crystals suitable for analysis by X-ray diffraction could be grown, complexes with perdeuterated ligand substituents were often used. The bis(l-oxo)dicopper cores in these structures are quite congruent despite differences in the nature of the supporting N-donor ligands. Thus, the Ê ) are short compared to typical Cu(II)AO CuAO distances (average = 1.81 A Ê ); bond-valence sum analyses suggested that the short distances (1.9±2.0 A CuAO and AN distances in these compounds are indicative of a Cu(III) Ê ) indicate the oxidation level [68]. Long O  O distances (average = 2.327 A

Fig. 13. X-ray structure of the bis(l-oxo) complex cation [{Cu(d21-Bn3TACN)}2(l-O)2]2+ [64].

Ê Selected bond lengths are given in A

196

A.G. Blackman á W.B. Tolman

absence of a bond between these atoms and the Cu  Cu separation Ê ) is much shorter than those found in other structurally (average = 2.773 A characterized binuclear copper-peroxo and superoxo species (Table 1). Overall, the [Cu2(l-O)2]2+ unit is structurally similar to other such cores with other metals (e.g., Fe [69], Mn [70], Ni, and Co [71]). The core is supported by tridentate N-donor macrocycles in [{Cu(d21-Bn3TACN)}2(l-O)2] (SbF6)2 á 7(CH3)2CO á 2CH3CN (Fig. 13) [64] and [Cu2(d28-i-Pr4DTNE)(lm-O)2] (SbF6)2 á 3CH2Cl2 (Fig. 14) [65], but different relative orientations are adopted by the square pyramidal metal ions; in the Bn3TACN case, the axial N-donors are disposed in an anti conformation, whereas in the i-Pr4DTNE instance they are forced by the ethylene tether to be syn. In both structures, close contacts between ligand substituent CAD bonds implicate CuAO  D bonding interactions that may lie along the decomposition pathways traversed by these bis(l-oxo) complexes (see below). In the structurally characterized complex [{Cu(LME)}2(l-O)2](CF3SO3)2 á 4CH2Cl2 (Fig. 15) [67] and analogous complexes containing the peralkylated cyclohexanediamine ligands LTM and LTE, bidentate N-donors lead to square planar metal ion geometries. Interestingly, use of the more ¯exible ligand LTMPD gave a bis(l-oxo) dimer that exhibited very different reactivity towards oxidizable substrates. EXAFS data suggested that this complex contained 5-coordinate copper centers [72]. Spectroscopic studies have revealed important signatures of the bis(loxo)dicopper core in these and related compounds [14, 66, 68], which in conjunction with theoretical calculations [73±77] have enabled the delineation

Fig. 14. X-ray structure of the bis(l-oxo) complex cation [Cu2(d28-i-Pr4DTNE)(l-O)2]2+

Ê [65]. Selected bond lengths are given in A

Fig. 15. X-ray structure of the bis(l-oxo) complex cation [{Cu(LME)}2(l-O)2]2+ [67]. Selected

Ê bond lengths are given in A

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

197

of key aspects of its electronic structure (see Note Added in Proof). A pair of intense optical absorbance features at 320 and 430 nm with e = 12,000± 16,000 M)1 cm)1 involve charge-transfer (presumably oxo ® Cu(III)) within the core, quite different from the pattern observed for other peroxo- or superoxo-copper compounds. Consistent with the UV/vis assignment, resonance Raman spectra obtained using excitation wavelengths within the 430 nm absorption band exhibited a strongly enhanced feature at 600 cm)1 that shifted by 23±29 cm)1 upon 18O-substitution. Detailed studies of the Raman spectra of a series of complexes with differing supporting ligands and isotopic compositions, as well as ab initio calculations on model systems, support assignment of this feature to a symmetric vibration of the Cu2O2 rhomb [78]. The bis(l-oxo)dicopper compounds are uniformly EPR silent and exhibit 1H NMR signals in the 0±10 ppm region, clearly re¯ecting their diamagnetism. On the basis of the geometric parameters and ab initio calculations, a bis(l-oxo)dicopper(III) valence bond description was suggested, with the caveat that the system is highly covalent due to strong bonding interactions between O p and Cu d orbitals in the HOMO (Fig. 16). A Cu X-ray absorption K-edge study provided experimental support for this notion [66]. The bis(l-oxo)dicopper(III) and previously discussed (l-g2:g2-peroxo) dicopper(II) cores are thus isomers potentially related by OAO bond making and breaking coupled with electron transfer to and from the copper ions, respectively. Experimental con®rmation of the interconversion between the two isomeric cores was accomplished for the system in which i-Pr3TACN was the supporting ligand [64]. It was subsequently shown that oxygenation of [Cu(i-Pr3TACN)(CH3CN)]+ yielded either or both isomers depending on the solvent, counterion, or temperature. For example, resonance Raman and UV/ vis spectroscopy showed that in CH2Cl2 the peroxo form predominated, but in concentrated THF solutions (>1 mmol l)1) the bis(l-oxo) core was generated instead. Interestingly, the cores could be interconverted in solvent mixing experiments; addition of THF (50-fold excess v/v) to a concentrated (1.2 mmol l)1) CH2Cl2 solution of the peroxo complex yielded the bis(l-oxo)

Fig. 16. The HOMO of anti [{(NH3)3Cu}2(l-O)2]2+ calculated at the RHF/STO-3G* level [68]

198

A.G. Blackman á W.B. Tolman

form, and the opposite protocol reversed the process. Rapid equilibration was initially implicated by stopped-¯ow kinetics investigations of the oxygenation reaction in acetone solution, where spectroscopy indicated the presence of a mixture of both cores. More recently, studies of dilute solutions of the system in CH2Cl2, THF, or mixtures of the two supported equilibration of the cores in these solvents as well, and revealed how the equilibrium behaved as a function of the solvent and temperature [79]. The combined data are consistent with closely similar thermodynamic stabilities for the isomeric [Cu2(l-O)2]2+ and [Cu2(l-g2:g2-O2)]2+ cores supported by i-Pr3TACN, as well as a low barrier for their interconversion. These conclusions have been rationalized by calculations at various levels of theory [73±77]. The reversible isomerization of the [Cu2(l-O)2]2+ and [Cu2(l-g2:g2-O2)]2+units has obvious consequences in the biological production and processing of dioxygen (see below). Using the information gleaned from the combined structural/spectroscopic studies of the crystallographically characterized bis(l-oxo)dicopper complexes in Figs. 13±15 as a basis, additional examples of such species have been studied. Oxygenation of the Cu(I) complex of the binucleating xylyl-based ligand m-XYLi-Pr4 resulted in the formation of both intra- and intermolecular products, with the product distribution depending on the solution concentration (Scheme 1). In dilute solution (80%). The crystal structure of the oxygenation product ``[{Cu(MePY2)}2O2](BArF)2'' showed metric Ê , Cu  Cu = 3.45 A Ê ) which parameters for the dicopper core (OAO = 1.67 A were intermediate between those values typical of the bis(l-oxo) and l-g2:g2peroxo cores, suggesting that the crystalline material was in fact a solid solution of complexes containing both cores. Such behavior has also been observed in crystals isolated from oxygenation mixtures of [Cu(i-Pr3TACN) (MeCN)]+ in CH2Cl2 [14]. 2.3 Trinuclear Complexes

The copper proteins ascorbate oxidase, ceruloplasmin, and laccase, which carry out the 4-electron reduction of O2 to H2O, have all been found to contain a trinuclear copper center at the active site [16] and attempts have been made to prepare tricopper-dioxygen species as models for these proteins. The synthesis of the trinuclear model complex [(LTM)3Cu3O2](CF3SO3)3 was reported in 1996 from the reaction of the Cu(I) precursor [(LTM)Cu(MeCN)]+

200

A.G. Blackman á W.B. Tolman

with O2 at )80°C in CH2Cl2 [87]. Manometry showed a 3:1 Cu/O2 ratio for this reaction and the dark brown product displayed intense bands in the UV/vis spectrum at 290 nm and 355 nm and less intense bands at 480 nm and 620 nm. No resonance Raman data for the complex were reported. The structure of the product was con®rmed by X-ray crystallography and a diagram of the cation is shown in Fig. 17. The structure consists of three copper ions, each in an approximately square-planar environment, coordinated to two nitrogens from the chelating Ê ) shows that the ligand and two oxygen atoms. The O  O distance (2.37 A OAO bond of dioxygen has been cleaved and that the structure is best formulated as containing two l3-oxo bridges. A Cu(II, II, III) formulation of this complex was suggested by the CuAO bond lengths within the trimer, with Ê and 1.98(1) A Ê ) being longer than the two sets of CuAO distances (2.01(1) A Ê ). This was also supported by subsequent magnetism and MCD third (1.83(1) A data, and molecular orbital calculation results, which also showed that the Cu(II) ions were ferromagnetically coupled [88]. The synthesis of this complex again shows that the Cu(III) oxidation state can be accessed simply with the use of molecular oxygen as oxidant. 2.4 Tetranuclear Complexes

A remarkable room-temperature stable tetranuclear Cu(II) complex containing a l4-bridging peroxide ligand was isolated as dark green crystals from the reaction of Cu(ClO4)2 á 6H2O with 4-methyl-2,6-bis(pyrrolidinomethyl) phenol (HL2) in MeOH in the presence of triethylamine and 3,5-di-tertbutylcatechol [89]. The crystal structure of [Cu4(L2)2(O2)(OMe)2(ClO4)]ClO4 á MeOH (Fig. 18) shows the four copper ions to be in square-pyramidal environments, each coordinated to a phenoxide, perchlorate, and methoxide oxygen atom, and a pyrrolidine nitrogen atom. Coordination about each

Fig. 17. X-ray structure of the trinuclear complex cation [(LTM)3Cu3O2]3+ [87]. Selected

Ê bond lengths are given in A

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

201

Fig. 18. X-ray structure of the tetranuclear complex cation [Cu4(L2)2(O2)(OMe)2(ClO4)]+

Ê [89]. Selected bond lengths are given in A

Ê) copper ion is completed by a l4-bridging peroxo ligand (OAO = 1.453(4) A which lies above the plane of the copper ions. This l4-bridging peroxo structural motif has also been observed in multinuclear complexes of Mo [90] and Fe [91]. The CuAO bond lengths involving the peroxo ligand are Ê and 1.939(2) A Ê. 1.961(2) A A subsequent paper described the synthesis of two further tetranuclear peroxo complexes using the piperidine- and morpholine-appended phenol ligands HL3 and HL4, and reported spectroscopic data on all three complexes [92]. The complexes exhibited similar UV/vis spectra, with bands around 285 nm, 390 nm, and 580 nm. These were assigned to ligand transitions, a combination of phenolate ® Cu(II) and peroxo ® Cu(II) charge transfers, and peroxo ® Cu(II) charge transfer transitions, respectively. The resonance Raman spectra of the three complexes showed bands assigned to m(OAO) at 878 cm)1, 898 cm)1, and 888 cm)1. Magnetic data collected on the structurally characterized complex suggested very strong antiferromagnetic coupling between the copper ions.

3 Reactivity of Oxygenase Model Complexes The oxygenation reactions mediated by such copper proteins as tyrosinase and dopamine b-hydroxylase involve incorporation of an O2-derived oxygen atom into an aromatic or aliphatic substrate [15, 16]. Such reactions necessarily involve substrate CAH activation and are generally proposed to occur via attack by an activated oxygen species. While oxidations of organic compounds by copper salts are well known [93], detailed mechanistic information such as that which would be obtained by studies of well-characterized copper-oxygen species is needed in order to understand and control selectivities [11].

202

A.G. Blackman á W.B. Tolman

Progress toward those goals is summarized below, with speci®c emphasis on recent results involving oxidations of ligand hydrocarbyl units during decompositions of the aforementioned copper-oxygen complexes. 3.1 Aromatic Oxygenation

Numerous examples exist in the literature of both endogenous and exogenous aromatic hydroxylation reactions mediated by Cu(I) complexes on reaction with O2. However, in many of these examples, the attempted spectroscopic characterization of the Cu/O2 intermediates has been thwarted by their high reactivity and/or low concentration in the reaction mixtures [94±96], and in only a few cases have the intermediates been well characterized. Hydroxylation of the benzene ring of the xylyl-based bridging ligand XYL-H was observed on low-temperature oxygenation of the dicopper(I) complex of this ligand and an intermediate having an absorbance maximum at 435 nm was observed spectroscopically [97]. Subsequent studies showed that the incorporated oxygen atom derived from molecular oxygen (Scheme 2) [98]. Detailed kinetic and thermodynamic studies suggested that the mechanism of the hydroxylation reaction was akin to that of electrophilic aromatic substitution, in that electron-donating substituents on the benzene ring gave increased rates of reaction, and that no deuterium isotope effect was observed. Operation of a radical mechanism was precluded by the observation that addition of radical traps to the reaction mixture did not affect the rate of reaction and it was assumed that a bent ``butter¯y'' l-g2:g2-peroxo species effected the hydroxylation reaction. However, the discovery that the l-g2:g2peroxo and bis(l-oxo) cores can interconvert [64] prompted a reinvestigation

Scheme 2. The reaction of [Cu2(XYL-H)]2+ with O2 [98]

Copper-Dioxygen and Copper-Oxo Species Relevant to Copper Oxygenases and Oxidases

203

of this system with a view to attempting to distinguish whether a l-g2:g2-peroxo or bis(l-oxo) species was responsible for the hydroxylation reaction [59]. To this end, the kinetics of reaction of the Cu(I) complex of the nitro-substituted xylyl ligand NO2-XYL were investigated. In this system, resonance Raman data suggested a maximum bis(l-oxo):l-g2:g2-peroxo ratio of 0.0013 which would require the bis(l-oxo) species to be 1000 times more reactive than the l-g2:g2-peroxo species to explain the observed kinetics. The authors contended that this reactivity difference was not supported by MO calculations and hence they proposed that the l-g2:g2-peroxo species reacted directly to give the hydroxylated product. As mentioned above (Sect. 2.2.4), oxygenation of the 3-coordinate Cu(I) complex [Cu(PhPyNEt2)(MeCN)](SbF6) in acetone or THF solution at )70 °C gave an EPR-silent yellow species that exhibited spectral features typical of a bis(l-oxo) core (UV/vis:kmax 404 nm, resonance Raman:607 cm)1, shifted to 580 cm)1 on substitution with 18O2) [81]. Hydroxylation at the ortho position of the phenyl substituent was observed when the oxygenated solution was allowed to warm to room temperature, with the yield for this process being 70% of theoretical. Identical behavior was observed when the oxygenation was carried out at room temperature, and 18O labeling con®rmed the incorporated oxygen atom derived from molecular O2 (Scheme 3). No isotope effect was observed on deuteration of the phenyl ring, arguing against CAH bond breaking being rate-determining. The rate of decay of the yellow species was found to be dependent on the nature of substituents on the phenyl group, with electron-withdrawing substituents giving slower rates and electron-donating substituents increasing the rate of reaction to the point where the intermediate yellow species could not be observed. Although the amount of l-g2:g2-peroxo species present in the reaction mixture was suggested to be very small (75%. Acetonitrile as solvent leads to the fastest reactions although the work-up is reported to be easier in CH2Cl2 [78]. Silyl enol ethers are oxidized to a-hydroxy ketones by MTO/H2O2 with successive disilylation with KF (Eq. 4) [79]. However, for this reaction to be successful, acetic acid and pyridine have to be present. It is assumed that pyridine prevents the formation of hydrolysis products while acetic acid is necessary to prevent base-induced MTO decomposition. Yields are usually >90%. In the case of conjugated systems, the yields are signi®cantly lower:

…4†

Rhenium-Oxo and Rhenium-Peroxo Complexes in Catalytic Oxidations

225

4.5 Oxidation of Alkynes

Internal alkynes yield carboxylic acids and a-diketones when oxidized with the MTO/H2O2 system (Eq. 5) [80]. Rearrangement products were observed only for aliphatic alkynes. Terminal alkynes give carboxylic acids, derivatives thereof, and a-keto acids as the major products. The yields of these products vary with the solvent used [80]. The mechanisms leading to the products observed remain highly speculative to date. Several possibilities are still under discussion.

…5†

4.6 Oxidation of Cyclic b-Diketones

The MTO/H2O2 system furthermore catalyzes the oxidation of cyclic b-diketones to carboxylic acids (Eq. 6) [81]. Conversions are usually above 85% and the product selectivity is nearly quantitative. The reaction is performed in a 1:1 water acetonitrile solution at room temperature. It has been assumed that enolic forms which exist in solution are ®rst epoxidized. After a rearrangement step the CAC bond is cleaved and an oxygen inserted. Then an a-diketone intermediate forms which is ®nally oxidized to the carboxylic acid [81].

…6†

4.7 Oxidation of Sulfur Compounds

Organic sul®des can be oxidized to the corresponding sulfoxides by hydrogen peroxide in the presence of MTO (Eq. 7) [82±84]. Both complexes 2 and 3 seem to be active in this reaction but kinetic results indicate that 2 might be more active than 3. The kinetic results also point to a mechanism that involves the nucleophilic attack of the sulfur atom on a coordinated peroxide oxygen since electron donating substituents have accelerating effects. Using ethanol as solvent, the MTO/H2O2 system can be used to oxidize dialkyl, diaryl and alkyl aryl sul®des to sulfoxides (R2S:H2O2 = 1:1.1) or sulfones (R2S:H2O2 = 1:2.2) with excellent yields and selectivity even in the presence of oxidatively

226

F.E. KuÈhn á W.A. Herrmann

sensitive functions on the sul®de side chain [82±89]; increased sulfone yields can also be achieved by higher catalyst concentrations. Functional groups in the side chain of the sul®de, e.g., carbon double bonds, are not affected under the reaction conditions: the sulfur atom is selectively oxidized [82]. MTO can also be used in the oxidation of sul®des with the water free system ureahydrogen peroxide as oxidant in acetonitrile. Partial enantioselective oxidation of methylphenyl sul®de was observed with MTO in the presence of a (+)camphor derived pyrazine carboxylic acid [83]. Trichloro oxorhenium phosphane complexes, e.g., mer,trans-Re(O)Cl3(PPh3)2, also catalyze the oxidation of sul®des to sulfoxides. These complexes do not so easily overoxidize the sul®des to sulfones as does the MTO-containing system. …7†

The mild selectivity of MTO/H2O2 has also been demonstrated in the oxidation of thioether Fischer carbene complexes to sulfoxides (Eq. 8)[85]:

…8†

Oxidation of thiophene and its derivatives has been achieved with this system. However, the rate constants are two to four orders of magnitude below those reported for the ``aliphatic'' sul®des where the S atom is not part of an aromatic heterocyclic ring [86]. The delocalization of electrons from the sulfur into the aromatic ring causes a decreased reactivity of the sulfur towards the electron-de®cient peroxide group of the metal center. MTO is able to catalyze the transformation of thiols to disul®des with DMSO (Eq. 9). The mechanism of this reaction is not completely clear. It is assumed that Re(V) intermediates may be involved [88]. …9†

Re(V) catalysts, e.g., Re(O)Cl3(PPh3)2, catalyze oxygen scrambling between sulfoxides (most preferably diphenyl sulfoxide) and sul®des (Eq. 10). Nearly quantitative yields are reached and the catalytic system is comparatively stable against moisture and oxygen [89]. …10†

Rhenium-Oxo and Rhenium-Peroxo Complexes in Catalytic Oxidations

227

Oxidation studies of coordinated thiolates were carried out on the model Co(III) complex [(en)2Co(SCH2CH2NH2)]2+. The thiolato complex is ®rst oxidized to a sulfenato complex, [(en)2Co(S(O)CH2CH2NH2)]2+, which is then more slowly oxidized to the sul®nato complex [(en)2Co(S(O)2CH2CH2NH2)]2+. The second step is ca. 1500 times slower than the ®rst. From these kinetic examinations, 2 seems to be the catalytically active species which is attacked by the nucleophilic S atom [87]. 4.8 Oxidation of Anilines and Amines

The MTO/H2O2 system also catalyzes the oxidation of anilines (Eq. 11). The major product of this oxidation is nitrosobenzene. For 4-substituted N,Ndimethylanilines, the N-oxide is the only oxidation product. Electronwithdrawing substituents inhibit the reaction. …11†

Kinetic results suggest that both compounds 2 and 3 are involved in the oxidation process [90]. It has been suggested that the rate-controlling step is the nucleophilic attack of the nitrogen lone pair electrons of the anilines on a peroxidic oxygen of the catalyst. Electron-donating groups attached to the nitrogen atom of aniline increase the rate constant. In the case of ArNH2 derivatives, the oxidation proceeds ca. 50 times faster than without catalyst [90]. In general, the reactions are facile and give high yields at or below room temperature [91]. Furthermore, a broad variety of aromatic and aliphatic secondary amines can be oxidized to the corresponding N-oxides [92±95]. The amines are converted to the corresponding hydroxylamines before conversion to the nitrones in very good yields [92±95]. The hydroxylamine formation is rate controlling [93]. Both H2O2 and the urea-hydrogen peroxide complex can be used together with MTO. Benzylamines are selectively oxidized to oximes. Primary amines are oxidized to nitro compounds by MTO catalysis (Eq. 12) [96, 97]: …12†

It is noteworthy in this context that the N-oxides also form adducts with MTO [56]. While aliphatic N-oxides form temperature-sensitive adducts with MTO and are inactive in the catalytic epoxidation of ole®ns, aromatic N-oxide adducts of MTO are both stable and catalytically active. In some cases their selectivity in ole®n epoxidation is even higher than with N-base adducts of MTO (see Sect. 4.1).

228

F.E. KuÈhn á W.A. Herrmann

4.9 Oxidative Conversion of N,N-Dimethylhydrazones to Nitriles

N,N-Dimethylhydrazones of aldehydes react with hydrogen peroxide in the presence of catalytic amounts of MTO to give in high yield the corresponding nitriles. Two different sets of conditions have been applied to perform this transformation. The reaction is either carried out adding dropwise the N,Ndimethylhydrazone in ethanol to a solution of 1.5% of MTO and 2 eq. of 35% H2O2 in EtOH at )50 °C [98] or in an acetonitrile-acetic acid-pyridine mixture (94.5:5:0.5) at room temperature with a 30% solution of H2O2 in water in the presence of MTO [99]. The pyridine is applied to suppress hydrolysis and the acetic acid is used to prevent deactivation of MTO by the basic hydrazones, according to the authors. The nitrile yields are reported in most cases considerably higher then 90%. The reaction starts with the oxidation of the N,N-dimethylhydrazone, and proceeds presumably through an intermediate shown in Eq. (13) which undergoes a Cope-type elimination to yield the nitrile [99, 100]:

…13†

4.10 Oxidation of Phosphines, Arsines, and Stibines

Tertiary phosphines, triaryl arsines, and triaryl stibines are converted to their oxides, R3EO (E = P, As, Sb) by MTO/H2O2. Kinetic studies lead to the assumption that 2 and 3 have similar catalytic activities in all cases. The kinetic data support a mechanism involving nucleophilic attack of the substrate at the rhenium peroxides. The proposed catalytic cycle is given in Scheme 4 [101]. In the absence of peroxides MTO also catalyzes the oxidation of tertiary phosphines to phosphine oxides [102, 103]. The oxygen sources in this case are sulfoxides which are deoxygenated. It is assumed that MTO forms a sulfoxide adduct ®rst which is then deoxygenated by triphenylphosphane under formation of MTO and sul®de. Another mechanism, suggesting reduction of MTO to methyldioxorhenium by the phosphine and reoxidation by the sulfoxide, seems to be less likely. The oxygen transfer from sulfoxides to phosphines is also catalyzed by several Re(V) species, e.g., mer,transRe(O)Cl3(PPh3)2 and fac,cis-Re(O)Cl3(CNCMe3)2 [104±106]. Mechanistic studies concerning these transfer reactions have been performed, showing that both ligand-centered and metal-centered oxygen-transfer reactions occur depending on the particular Re(V) catalyst used [105, 106].

Rhenium-Oxo and Rhenium-Peroxo Complexes in Catalytic Oxidations

229

Scheme 4.

4.11 Oxidation of CAH Bonds

MTO/H2O2 also catalyzes the insertion of oxygen into a variety of activated and inactivated CAH bonds with yields varying from good to excellent. Alcohols or ketones are formed as shown in Eqs. (14) and (15). In the case of tertiary substrates, alcohols are obtained as the products. Suitable substrates proved that the reaction is stereo speci®c with retention of the con®guration. Alcohols, e.g., ethanol and t-butanol, or acetonitrile are used as solvents. The reaction temperatures range between 40 °C and 80 °C and nearly quantitative yields have been obtained in several cases. However, the reaction times are generally longer than those used for most epoxidations, ranging between 10 h and 72 h [107]. The reaction can be accelerated by the addition of pyrazine-2-carboxylic acid, and the total yield is also increased [108]. …14†

230

F.E. KuÈhn á W.A. Herrmann

…15†

4.12 Oxidation of SiAH Bonds

The catalyst system MTO/H2O2 also catalyzes oxygen atom insertion into SiAH bonds. Silanols and siloxanes are formed as products, the latter being the main products [107, 109]. When the urea/hydrogen peroxide adduct (UHP) is used as oxygen source instead of aqueous H2O2, MTO catalyzes the oxidation of silanes to silanols in high conversions (>95% within 18 h at room temperature) and excellent selectivities (usually 90%) in favor of the silanol (only tiny amounts of disiloxane are formed, if at all) (Eq. 16). It is assumed that the SiAH oxidation takes place in the helical urea channels. The urea matrix serves as a host for the silane substrate, the H2O2 oxygen source, and the MTO metal catalyst as guest. In the con®ned environment of these channels the catalyst is stabilized against decomposition while the condensation of silanol to disiloxane is avoided for steric reasons [109]. Other matrices than urea have been examined too, namely amylase and SiO2, but neither match the good results obtained with urea. In both cases the conversions are lower and the amount of disiloxane obtained is considerably higher [109].

…16†

4.13 Oxidation of Aromatic Compounds

Arenes are oxidized to p-benzoquinones by hydrogen peroxide with MTO as catalyst [110±118]. Noteworthy is the high regioselectivity, most notably in the industrially interesting synthesis of vitamin K3 (see Eq. 17): …17†

Since water is an inhibitor, concentrated (85 wt%) H2O2 is preferred. Alternatively, commercially available 35% H2O2 in acetic anhydride can be employed; a considerable regioselectivity is obtained with this system. The conversion is higher for electron-rich arenes (nearly 100%) and selectivities of more than 85% have been reached [110]. Biphenylene can be oxidized

Rhenium-Oxo and Rhenium-Peroxo Complexes in Catalytic Oxidations

231

with the MTO system in chloroform affording an o-quinone product (83% conversion) [111]. Hydroxy substituted arenes can be oxidized by aqueous hydrogen peroxide (85 wt%) in acetic acid to afford the corresponding pquinonones in isolated yields of up to 80% [114]. It has been shown that using a mixture of acetic acid and acetic anhydride further improves the product yield [117, 118]. Instead of acetic acid, HBF4 in EtOH can be used too [114, 115]. Anisol was also found to undergo selective oxidation with the MTO/H2O2 system to yield o- and p-methoxyphenols. There is no need to use a solvent in this case. Bidentate Lewis base adducts of halogeno rhenium(VII) oxides also catalyze the arene oxidation in the presence of H2O2 [116].

…18†

Concerning the mechanism of the arene oxidation it is assumed that the ®rst reaction step proceeds according to Eq. (18) to form a phenol. The phenols and methoxy benzenes, being more electron-rich than arenes, are more reactive with the peroxorhenium compounds. Therefore, the subsequent steps in the oxidation occur more rapidly than the ®rst [118]. Benzaldehydes with hydroxyl or methoxy substituents in ortho or para positions are oxidized to the corresponding phenols (carboxylic acids are formed as by product) in good yields [113] The yield is temperature and solvent dependent (Eq. 19):

…19†

4.14 Baeyer-Villiger Oxidation and Dakins Reaction

It has been shown that 3 also acts as an active species in the Baeyer-Villiger oxidation of ketones (Eq. 20) and in the Dakins reaction [119, 120]. …20†

232

F.E. KuÈhn á W.A. Herrmann

It is somewhat surprising that the MTO/H2O2 system presents this activity since these oxidations involve nucleophilic attack at the carbonyl group which contrasts with all the preceding examples where the substrates attacked the electrophilic Re-peroxo complexes, e.g., in the ole®n epoxidation. Nevertheless, 3 reacts stoichiometrically with cyclobutanone in the absence of H2O2. This reversed behavior may be due to substrate binding to rhenium. The unsymmetric geometry of compound 3a, displaying a polarity within the peroxo ligands might also be responsible for the observed behavior. The reaction was found to be strongly solvent dependent by means of the mechanistic probe for oxygen transfer reactions, thiantrene-5-oxide [81, 119]. Donor solvents such as acetonitrile seem to enhance the nucleophilicity of the peroxo groups. It has been suggested that the double bond of the enol form (the major tautomer) attacks a peroxo oxygen of the peroxorhenium complex 2 or 3. This reaction affords a 2-hydroxy-1,3-dicarbonyl intermediate which can be detected by 1H-NMR. This hydroxy intermediate is susceptible to cleavage via Baeyer-Villiger oxidation to yield carboxylic acids as ®nal products. Low H2O2 concentrations are suf®cient and no H2O2 decomposition is observed at temperatures up to 70 °C. This is an advantage of the catalytic MTO/H2O2 system over the known transition-metal Baeyer-Villiger catalysts containing V, Mo, Mn, or Os. However, the nucleophilic character of the peroxidic atoms in 3 is not as pronounced as in Pt or Ir peroxo complexes that react with CO2 or SO2 to give isolable cycloaddition products [119]. In the case of 3 turnover frequencies of 18000 [mol/mol cat. per hour] are obtained for cyclobutanone, but in other cases turnover numbers of ca. 100 are usual [120]. Cycloketones can be converted into lactones even below room temperature (15 °C) by diluted hydrogen peroxide (10 wt%). 4.15 Oxidation of Halide Ions

Another application of the MTO/H2O2 system is the catalytic oxidation of chloride and bromide ions in acidic aqueous solutions. The chloride oxidation steps are three to four orders of magnitude slower than the corresponding bromine oxidation steps. Both compounds 2 and 3 have been shown to be active catalysts in these processes. In both cases the catalyzed reactions were about 105 times faster than the uncatalyzed reactions under similar conditions. In a ®rst step HOX is formed, then HOX reacts with X) to form X2. When H2O2 is used in excess the reaction yields O2 [121, 122]. 4.16 Oxidation of Metal Carbonyls

MTO and some of its derivatives, e.g., CpReO3 and EtReO3, catalyze the oxidation of metal carbonyls to metal oxides with H2O2 [13, 123, 124]. This reaction runs at room temperature and yields of up to 90% are obtained. However, only organometal carbonyls with oxidation resistant organic groups can be oxidized, e.g., (pentamethylcyclopentadienyl) tricarbonyl rhenium(I)

Rhenium-Oxo and Rhenium-Peroxo Complexes in Catalytic Oxidations

233

(Eq. 21) [13]. In all other cases the organic ligand is also oxidized, leading to decomposition of the product complex [124].

…21†

5 Conclusions Rhenium(VII) oxides have found numerous applications in catalysis during the recent years. MTO, a water soluble, air and temperature stable metal alkyl has become the most versatile oxidation catalyst in organic chemistry. Its outstanding reactivity in oxidation catalysis is due to the highly Lewis acidic and sterically unsaturated rhenium(VII) center and the thermally and chemically very stable rhenium-carbon bond. Inorganic rhenium oxides, e.g., Re2O7, and their peroxo complexes are far less versatile in watercontaining systems due to their pronounced moisture sensitivity, but can be applied in water free oxidation systems, e.g., with (Me3SiO)2. However, in spite of a rich body of MTO chemistry and applications available now, this area is far from being mature. This is clearly seen by the still improving number of papers published each year on new applications of MTO. Several derivatives of MTO, e.g., CpReO3, are also stable enough to be used as catalysts but have not been examined in detail concerning their catalytic abilities. We particularly believe that organometal oxides and derivatives in rhenium's neighborhood of the periodic table will still improve the catalytic strength of this class of compounds.

6 References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.

Rouschias G (1974) Chem Rev 74: 531 Herrmann WA, KuÈhn FE (1997) Acc Chem Res 30: 169 Warwel S, RuÈsch M, Sojka M (1991) J Chem Soc Chem Commun 1587 Noddack J, Noddack W (1929) Z Anorg Allg Chem 181: 1 Smith WT, Line LE, Bell WA (1952) J Am Chem Soc 74: 4964 Wells AF (1984) Structural inorganic chemistry, 5th edn. Clarendon, Oxford, pp 509± 550 Krebs B, MuÈller A, Beyer HH (1968) J Chem Soc Chem Commun 263 Krebs B, MuÈller A (1968) Z Naturforsch 23b: 415 Beyer H, Glemser O, Krebs B, Wagner G (1970) Z Anorg Allg Chem 376: 87 Beyer H, Glemser O, Krebs B (1968) Angew Chem Int Ed Engl 7: 295 Fischer D, Krebs B (1982) Z Anorg Allg Chem 491: 73 Krebs B, Fischer D, Paul G (1996) Z Anorg Allg Chem 622: 448 Herrmann WA, Correia JDG, KuÈhn FE, Artus GRJ, RomaÄo CC (1996) Chem J Eur 2: 168

234

F.E. KuÈhn á W.A. Herrmann

14. Kiprof P, Herrmann WA, KuÈhn FE, Scherer W, Kleine M, Elison M, Rypdal K, Volden HV, Gundersen S, Haaland A (1992) Bull Soc Chim Fr 129: 655 15. Herrmann WA, Roesky PW, KuÈhn FE, Elison M, Scherer W, Artus GRJ, RomaÄo CC, Lopes AD, Basset JM (1995) Inorg Chem 34: 4701 16. Mertis K, Wilkinson G (1976) J Chem Soc Dalton Trans 1488 17. Beattie IR, Jones PJ (1979) Inorg Chem 18: 2318 18. Herrmann WA, Kuchler JG, Felixberger JK, Herdtweck E, Wagner W (1988) Angew Chem Int Ed Engl 27: 394 19. Herrmann WA, KuÈhn FE, Fischer RW, Thiel WR, RomaÄo CC (1992) Inorg Chem 31: 4431 20. Herrmann WA, Thiel WR, KuÈhn FE, Fischer RW, Kleine M, Herdtweck E, Scherer W, Mink J (1993) Inorg Chem 32: 5188 21. Herrmann WA, Kratzer R, Fischer RW (1997) Angew Chem Int Ed Engl 36: 2652 22. Herrmann WA (1988) Angew Chem Int Ed Engl 27: 394 23. Herrmann WA (1995) J Organomet Chem 500: 149 24. KuÈhn FE, Fischer RW, Herrmann WA (1999) Chem i u Zeit 33: 192 25. Espenson JH (1995) Chemical kinetics and reaction mechanisms. McGraw-Hill, New York 26. RomaÄo CC, KuÈhn FE, Herrmann WA (1997) Chem Rev 97: 3197 27. Cusumaro JA (1992) CHEMTECH 482 28. Schmidt B (1997) J Prakt Chem 339: 439 29. Espenson JH, Abu-Omar MM (1997) ACS Adv Chem 253: 3507 30. Gable KP (1997) Adv Organomet Chem 41: 127 31. Brown SN, Mayer JM (1996) J Am Chem Soc 118: 12,119 32. Herrmann WA, Fischer RW, Marz DW (1991) Angew Chem Int Ed Engl 30: 1638 33. Yamazaki S, Espenson JH, Huston P (1993) Inorg Chem 32: 4683 34. Herrmann WA, Correia JDG, Artus GRJ, Fischer RW, RomaÄo CC (1996) J Organomet Chem 520: 139 35. Herrmann WA, Fischer RW, Scherer W, Rauch MU (1993) Angew Chem Int Ed Engl 32: 1157 36. Herrmann WA, Fischer RW, Rauch MU, Scherer W (1994) J Mol Catal 86: 243 37. Espenson JH (1999) Chem Commun 479 38. Gisdakis P, Antonczak S, KoÈstlmeier S, Herrmann WA, RoÈsch N (1998) Angew Chem Int Ed Engl 37: 2211 39. Abu-Omar MM, Hansen PJ, Espenson JH (1996) J Am Chem Soc 118: 4966 40. Laurenczy G, LukaÂcs F, Roulet R, Herrmann WA, Fischer RW (1996) Organometallics 15: 848 41. Espenson JH, Tan H, Mollah S, Houk RS, Eager MD (1998) Inorg Chem 37: 4621 42. Brittingham KA, Espenson JH (1999) Inorg Chem 38: 744 43. SchroÈder D, Fiedler A, Herrmann WA, Schwarz H (1995) Angew Chem Int Ed Engl 34: 2636 44. Pestovsky O, v. Eldik R, Huston P, Espenson JH (1995) J Chem Soc Dalton Trans 133 45. Hatzopoulos I, Brauer HD, Geisberger M, Herrmann WA (1996) J Organomet Chem 520: 201 46. Al-Ajlouni AM, Espenson JH (1995) J Am Chem Soc 117: 9243 47. Herrmann WA (1993) In: Herrmann WA (ed) Organic peroxygen chemistry, vol 164. Springer, Berlin Heidelberg New York, p 130 48. Boehlow TR, Spilling CD (1996) Tetrahedron Lett 37: 2717 49. Murray RW, Singh M, Williams BL, Moncrieff HM (1995) Tetrahedron Lett 36: 2437 50. Al-Ajlouni AM, Espenson JH (1996) J Org Chem 61: 3969 51. KuÈhn FE, Santos AM, Roesky PW, Herdtweck E, Scherer W, Gisdakis P, Yudanov IB, Di Valentin C, RoÈsch N (1999) Chem Eur J 5: 3603 52. Adam W, Mitchell CM (1996) Angew Chem Int Ed Engl 35: 533 53. Suzuki T, Iida K, Wada K, Kondo T, Mitsudo T (1999) Tetrahedron Lett 40: 2997 54. Murray RW, Iyanar K (1997) Tetrahedron Lett 38: 335

Rhenium-Oxo and Rhenium-Peroxo Complexes in Catalytic Oxidations

235

55. Herrmann WA, KuÈhn FE, Mattner MR, Artus GRJ, Geisberger M, Correia JDG (1997) J Organomet Chem 538: 203 56. Herrmann WA, Correia JDG, Rauch MU, Artus GRJ, KuÈhn FE (1997) J Mol Catal 118: 33 57. Rudolph J, Redding KL, Chiang JP, Sharpless KB (1997) J Am Chem Soc 119: 6189 58. Yudin AK, Sharpless KB (1997) J Am Chem Soc 119: 11,536 59. CopeÂret C, Adolfsson H, Sharpless KB (1997) Chem Commun 1565 60. Herrmann WA, Wojtczak WA, Artus GRJ, KuÈhn FE, Mattner MR (1997) Inorg Chem 35: 465 61. Herrmann WA, Ding H, Kratzer R, KuÈhn FE, Haider JJ, Fischer RW (1997) J Organomet Chem 97: 3197 62. Nakajima M, Sasaki Y, Iwamoto H, Hashimoto S (1998) Tetrahedron Lett 39: 87 63. Herrmann WA, Kratzer RM, Ding H, Thiel WR, Glas H (1998) J Organomet Chem 555: 293 64. GansaÈuer A (1997) Angew Chem Int Ed Engl 36: 2591 65. Rudler H, Gregorio JR, Denise B, BreÂgeault JM, Deloffre A (1998) J Mol Catal 133: 255 66. Wang WD, Espenson JH (1998) J Am Chem Soc 120: 11,335 67. Villa de PAL, De Vos DE, Montes de C C, Jacobs PA (1998) Tetrahedron Lett 39: 8521 68. Tan H, Espenson JH (1998) Inorg Chem 37: 467 69. Tetzlaff HR, Espenson JH (1999) Inorg Chem 38: 881 70. Adam W, Corma A, MartõÂnez A, Mitchell CM, Reddy TI, Renz M, Smerz AK (1997) J Mol Catal 117: 357 71. Zhu Z, Espenson JH (1996) J Org Chem 61: 324 72. Adam W, Mitchell CM, Saha-MoÈller CR (1999) Eur J Org Chem 785 73. Jacob J, Espenson JH, Jensen JH, Gordon MS (1998) Organometallics 17: 1835 74. Bellemin-Laponnaz S, LeNy JP, Dedieu A (1999) Chem Eur J 5: 57 75. Zauche TH, Espenson J (1998) Inorg Chem 37: 6827 76. Espenson JH, Zhu Z, Zauche TH (1999) J Org Chem 64: 1191 77. Herrmann WA, Zoller JP, Fischer RW (1999) J Organomet Chem 581: 404 78. Finlay J, McKervey A, Gunaratne HQN (1998) Tetrahedron Lett 5651 79. Stancovic S, Espenson JH (1998) J Org Chem 63: 4129 80. Zhu Z, Espenson JH (1995) J Org Chem 60: 7728 81. Abu-Omar MM, Espenson JH (1996) Organometallics 15: 3543 82. Yamazaki S (1996) Bull Chem Soc Jpn 69: 2955 83. Gunaratne HQN, McKervey MA, Feutren S, Finlay J, Boyd J (1998) Tetrahedron Lett 39: 5655 84. Adam W, Mitchell CM, Saha-MoÈller CR (1994) Tetrahedron 50: 13,121 85. Beddoes RL, Painter JE, Quayle P, Patel P (1996) Tetrahedron Lett 37: 9385 86. Brown KN, Espenson JH (1996) Inorg Chem 35: 7211 87. Huston P, Espenson JH, Bakac A (1993) Inorg Chem 32: 4517 88. Jacob J, Guzei IA, Espenson JH (1999) Inorg Chem 38: 1040 89. Arterbum JB, Nelson SL (1998) J Org Chem 61: 2260 90. Zhu Z, Espenson JH (1995) J Org Chem 60: 1326 91. Vassell KA, Espenson JH (1994) Inorg Chem 33: 5491 92. Goti A, Nannnelli L (1996) Tetrahdron Lett 37: 6025 93. Zauche TH, Espenson JH (1997) Inorg Chem 36: 5257 94. Murray RW, Iyanar K, Chen I, Wearing T (1996) J Org Chem 61: 8099 95. Yamazaki S (1997) Bull Chem Soc Jpn 70: 877 96. Murray RW, Iyanar K, Chen I, Wearing T (1996) Tetrahedron Lett 37: 809 97. Murray RW, Ianar K (1996) J Org Chem 61: 8099 98. Rudler H, Denise B (1998) Chem Commun 2145 99. Stancovic S, Espenson JH (1998) Chem Commun 1579 100. Zhu Z, Espenson JH (1997) J Mol Catal 121: 139 101. Abu-Omar MM, Espenson JH (1995) J Am Chem Soc 117: 272 102. Zhu Z, Espenson JH (1995) J Mol Catal 103: 87

236

F.E. KuÈhn á W.A. Herrmann

103. 104. 105. 106. 107. 108. 109. 110.

Abu-Omar MM, Appelman EH, Espenson JH (1996) Inorg Chem 35: 7751 Abu-Omar MM, Khan SI (1998) Inorg Chem 37: 4979 Bryan JC, Stenkamp RE, Tulip TH, Mayer JM (1987) Inorg Chem 26: 2283 Rybek WK, Zagiczek AJ (1992) Coord Chem 26: 79 Murray RW, Iyanar K, Chen L, Wearing JT (1995) Tetrahedron Lett 36: 6415 Schuchardt U, Mandelli D, Shul'pin GB (1996) Tetrahedron Lett 37: 6487 Adam W, Mitchell CM, Saha-MoÈller CR, Weichold O (1999) J Am Chem Soc 121: 2097 Adam W, Herrmann WA, Lin J, Saha-MoÈller CR, Fischer RW, Correia JDG (1994) Angew Chem Int Ed Engl 33: 2475 Adam W, Balci M, Kilic H (1998) J Org Chem 63: 8544 Karasevich EI, Nikitin AV, Rubailo VL (1994) Kinet Katal 35: 810 Yamazaki S (1995) Chem Lett 127 Adam W, Herrmann WA, Lin J, Saha-MoÈller CR (1994) J Org Chem 59: 8281 Adam W, Herrmann WA, Saha-MoÈller CR, Shimizu M (1995) J Mol Catal 97: 15 KuÈhn FE, Haider JJ, Herdtweck E, Herrmann WA, Lopes AD, Pillinger M, RomaÄo CC (1998) Inorg Chim Acta 279: 44 Herrmann WA, Haider JJ, Fischer RW (1999) J Mol Cat 270: 55 Jacob J, Espenson JH (1998) Inorg Chim Acta 270: 55 Herrmann WA, Fischer RW, Correia JDG (1994) J Mol Catal 94: 213 Fischer RW (1994) Ph D Thesis, Technical University Munich Espenson JH, Pestovsky O, Huston P, Staudt S (1994) J Am Chem Soc 116: 2869 Hansen PJ, Espenson JH (1995) Inorg Chem 34: 5389 Thiel WR, Fischer RW, Herrmann WA (1993) J Organomet Chem 459: C9 Herrman WA, Geisberger MR, KuÈhn FE, Artus GRJ, Herdtweck E (1997) Z Anorg Allg Chem 623: 1229

111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124.

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes Waldemar Adam, Catherine M. Mitchell, Chantu R. Saha-MoÈller, Oliver Weichold Institut fuÈr Organische Chemie der UniversitaÈt WuÈrzburg, Am Hubland, D-97074 WuÈrzburg, Germany E-mail: [email protected]; http://www-organik.chemie.uni-wuerzburg.de

Dioxiranes, MTO/H2O2, and MTO/UHP are becoming increasingly important as ef®cient and selective oxidants in organic chemistry and numerous publications have appeared on each of them. Here we present a comprehensive review, comparing the chemistry of these reagents with particular emphasis on the oxidation of p bonds, heteroatoms, and r bonds. We will discuss structural aspects and on this basis, will outline the reactivity and selectivity of these oxidants, along with the mechanistic aspects of the oxygen-transfer process. Keywords: Peroxo complexes, Dioxiranes, Oxygen transfer, Reactivity, Chemoselectivity,

Regioselectivity, Diastereoselectivity, Enantioselectivity

1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238

2

Structural Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242

3

Reaction Types and Reactivity . . . . . . . . . . . . . . . . . . . . . . . . . 243

3.1 3.2 3.3 3.4

Oxidation of p Bonds (Olefins, Alkynes, Oxidation of Heteroatoms (S, N) . . . . . Oxidation of r Bonds (CAH, SiAH) . . . Other Oxidation Types . . . . . . . . . . . .

4

Mechanistic Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . 261

4.1 4.2 4.3

p-Bond Oxidations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262 Heteroatom Oxidations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263 r-Bond Oxidations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264

5

Selectivities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265

5.1 5.2 5.3 5.4

Chemoselectivity . . Regioselectivity . . . Diastereoselectivity Enantioselectivity .

6

Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279

7

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

Arenes) . ....... ....... .......

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

244 248 256 260

265 270 274 277

Structure and Bonding, Vol. 97 Ó Springer Verlag Berlin Heidelberg 2000

238

W. Adam et al.

List of Abbreviations 1,2

A A con®g. convn Cp* Dec DMD Hept mal MTO NaY Np O pic quin SOSO SSO SSO2 t T TFA TFD TON UHP 1,3

1,2-allylic strain 1,3-allylic strain con®guration conversion pentamethylcyclopentadienyl decyl dimethyldioxirane heptyl malonyl methyltrioxorhenium Na-exchanged Y-zeolite with faujasite (FAU) structure naphthyl oxidant picolinyl quinoline thianthrene 5,10-dioxide thianthrene 5-oxide thianthrene 5,5-dioxide time temperature tri¯uoroacetone methyl(tri¯uoromethyl)dioxirane turnover number urea-hydrogen-peroxide adduct

1 Introduction Oxygen-transfer reactions are of increasing interest to both organic and inorganic chemists and have been the subject of an intensifying partnership [1]. Due to structural similarities between `organic' and `organometallic' oxidants, the study of common reactions and mechanisms, and also their differences, is essential for a better understanding of this burgeoning research ®eld, the goal being the development of new and more ef®cient oxygentransferring reagents. Of particular relevance for oxyfunctionalization reactions are those reagents which allow mild reaction conditions and lead to the desired product in both high yield and selectivity. Catalytic reaction conditions are desirable, but should not be enforced at the price of selectivity. A wide range of both organic and organometallic oxygen-transferring agents has been studied to date with respect to their reactivity and selectivity in oxyfunctionalizations. Different types of metal-based oxidants have been scrutinised, i.e., oxo, superoxo, peroxo, peroxy, perhydrate complexes (Fig. 1), and numerous reports have appeared on oxygen-transfer reactions with such oxidants. We will limit our review to the reactivity and selectivity of rhenium peroxo complexes derived from methyltrioxorhenium (MTO) and H2O2, but

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

239

Fig. 1. Types of metal complexes used for oxygen-transfer reactions

for convenience an overview of selected examples of oxygen-transfer reactions for other transition-metal g-peroxo complexes is also given in Table 1. Most of the examples use stoichiometric amounts of the complexes and disadvantages such as poor selectivity, double-bond cleavage, and cis/trans isomerization are common problems. Nonetheless, Table 1 demonstrates that peroxo complexes are known for most transition metals, which are capable of transferring oxygen. Hydrogen peroxide usually serves as the oxygen source for early transition metals and molecular oxygen for the late ones. Among these, the rhenium diperoxo complex CH3ReO(O2)2 á H2O stands out in view of its high reactivity and few side-reactions compared to the other peroxo complexes in Table 1. Indeed, none of the other peroxo complexes competes with the rhenium peroxo complex CH3ReO(O2)2 á H2O in terms of the broad scope of oxyfunctionalisation (cf. Scheme 1) and its catalytic ef®cacy. The structural similarity between metal peroxo complexes and the organic dioxiranes is made evident in Fig. 2, and the goal of this essay is, therefore, a comparison of the prototypes of these two classes of peroxides, namely the catalytic rhenium-based MTO/H2O2 oxidant (rhenium peroxo complexes have been proposed as the active oxygen-transferring agent [13]) and the stoichiometric dimethyldioxirane (DMD) and methyl(tri¯uoromethyl)dioxirane (TFD). We shall focus on the work from our research group, since we have for some time employed these peroxides in the selective synthesis of oxyfunctionalized products and in the mechanistic elucidation of the oxygentransfer processes on a comparative basis. After an initial brief discussion of the structural parameters of the rhenium peroxo species and of the most commonly utilized dioxiranes, a comparison of their reactivity and selectivity shall be made, and ®nally the mechanistic features of these oxygentransferring agents will be highlighted.

Fig. 2A,B. Structural similarity between: A metal-peroxo complexes; B dioxiranes

240

Table 1. Transition-metal peroxo complexes used in oxygen-transfer reactions

Oxygen source

Reaction type/preferred substrates

Drawbacks

Ti(O2)(OEt)2(H2O)2

H2O2

poor selectivity

Zr(O2)(mal)(quin)a VO(O2)(pic)(H2O)b2

H2O2 H2O2

± ± ± ± ± ±

Nb(Cp)2(O2) Cl CrO(O2)2L (L = pyridine, HMPA)

H2O2 H2O2

± oxidation of phosphines ± oxidation of alcohols to carbonyl compounds

MoO(O2)2(H2O)(HMPA)

H2O2

WO(O2)2(H2O)(HMPA)

H2O2

CH3ReO(O2)2(H2O) Ru(O2)Cl(NO)(PPh3)2 Co2(CN)4(PMe2Ph)5(O2) [Rh(O2)(AsMe2Ph)4](ClO4)

H2O2 O2 O2 O2

Ni(O2)(t-BuNC)2

O2

Pt(O2)(PPh3)2

O2

a b

± ± ± ± ± ± ± ± ± ± ± ± ± ±

Hmal = malonic acid; quin = quinoline. Hpic = picolinic acid.

epoxidation of cyclohexene hydroxylation of phenol oxidation of PPh3 epoxidation hydroxylation of arenes hydroxylation of alkanes

oxidation of hydroquinones to quinones sulfoxidation N-oxidation epoxidation oxidation of alcohols to carbonyl compounds phosphine oxidation epoxidation oxidation of alcohols to carbonyl compounds cf. Scheme 1 oxidation of phosphines oxidation of phosphines co-oxidation of terminal ole®ns to methyl ketones and PPh3 to OPPh3 oxidation of phosphines oxidation of isocyanides to isocyanates oxidation of phosphines

Stoichiometric or catalytic?

Refs. [2]

stoichiometric stoichiometric

[3] [4]

double-bond cleavage

stoichiometric stoichiometric

[5] [6]

double-bond cleavage

stoichiometric

[7]

stoichiometric

[1b]

catalytic stoichiometric stoichiometric stoichiometric

[8] [9] [10]

stoichiometric

[11]

stoichiometric

[12]

cis/trans isomerisation double-bond cleavage mixture of alcohols and ketones

W. Adam et al.

Peroxo complex

241

Scheme 1. Major oxidative transformations of MTO/H2O2 and dioxiranes

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

242

W. Adam et al.

2 Structural Aspects Although the dioxirane structure was postulated for the ®rst time in 1899 (as an intermediate in the Baeyer-Villiger oxidation of menthone by Caro¢s acid), convincing evidence for its existence appeared only in the late 1970s through to the mid 1980s by means of microwave and NMR spectroscopy [14]. The year 1985 marks the crowning achievement of the isolation of DMD as acetone solution [15], in 1994 the ®rst crystalline dioxirane, i.e., dimesityldioxirane [16], was prepared. Structural information of the rhenium diperoxo complexes CH3ReO(O2)2 á L (L=H2O, 4-tert-butylpyridine N-oxide or HMPA) was made available by X-ray crystallography of the adduct of CH3ReO(O2)2 á H2O with diglyme [17] and corroborated by NMR and IR spectroscopy [18]. For both the rhenium mono- and diperoxo complexes [19] and some dioxiranes [20], theoretical work con®rmed that the three-membered-ring peroxo-type structure is of lower energy than its isomeric acyclic form. Despite their geometrical similarities, the structural parameters (bond angles and lengths) of the rhenium peroxo functionality and of dioxiranes are distinct. The data in Table 2 reveal that the OAO bond lengths are comparable (147.4 and 151.6 pm). The MAO and CAO bond lengths, however, differ drastically since the MAO bond is ca. 30% longer than the CAO bond in the dioxiranes. This difference has consequences for the reactivity and selectivity of these oxidants, as will be discussed later. From the UV spectra it is evident that a peroxo functionality is present in the MTO/H2O2 and dioxirane Table 2. Spectral and structural data of rhenium peroxo complexes and dioxiranes

Peroxide

a b c d

M-O or C-O [pm]

OAO [pm]

UV/Vis [nm]

17 O-NMR [ppm]a

References

±

±

310

±

[17]

190.4; 191.8b

147.4b

364

363; 422

[17]

138.8c

151.6c

±

250

[14a,b]

141.7d

152.1d

335

302

[14b, 21]

±

±

347

297

[14c]

Values only given for the peroxo oxygen atoms. Determined by X-ray crystallography. Determined by microwave spectroscopy. Calculated by MP2/6-31G*.

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

243

oxidants. The absorption bands for both the mono- and diperoxo complexes (310 nm and 364 nm respectively) as well as for DMD (335 nm) and TFD (347 nm) are located in the near UV, which results in a yellow colour for these species. The 17O-NMR data (which are characteristic for mono- and diperoxo complexes) constitute further evidence for the peroxo ring in the rhenium oxidant [22].

3 Reaction Types and Reactivity Numerous oxidative transformations have been performed with both MTObased oxidants and dioxiranes (Scheme 1), which are known to be highly active oxygen-transferring species. These reactions include the oxidation of p bonds, r bonds and heteroatoms. Depending on the reaction conditions, high yields and selectivities have been achieved for most of these reaction types with both oxidants. Although MTO/ H2O2 has been claimed as one of the most active catalytic oxidation systems [18], compared to DMD, it is less reactive; however, DMD is used stoichiometrically and MTO catalytically. The choice of solvent and the appropriate additive may improve the reactivity and selectivity for the MTO catalyst. The MTO catalyst only ef®ciently activates hydrogen peroxide for oxidation reactions, but not other oxygen sources such as molecular oxygen, alkyl hydroperoxides, and alkyl peroxides [23]. Also, Me3SiOOSiMe3 has been employed, but presumably adventitious moisture hydrolyses the sensitive silyl peroxide into H2O2 and Me3SiOSiMe3 [24] and, thus, the real oxygen source is again H2O2. For the preparation of dioxiranes, potassium persulfate (commercial names Caroate, Curox, or Oxone) has proved to be the only practical oxygen donor; even then, most of the persulfate is decomposed into molecular oxygen and the yield of the isolated dioxirane is low (ca. 5% for DMD) [25]. In the in-situ generation [26] the yield of oxidation product may be quite high, especially if the substrate is very reactive. Rigorous pH control is necessary, as the dioxirane is otherwise decomposed to the ketone and molecular oxygen [27]. Although dioxiranes are usually used stoichiometrically, catalytic variants by in-situ generation have recently been developed for asymmetric oxidations [28]. Although the oxidants MTO/H2O2 and DMD show similar reaction behaviour, a number of differences are, nonetheless, observed: Some oxidations are only possible with MTO/H2O2 or with DMD; also, differences in their oxidative ef®cacy have been noted. A major difference in their reaction behaviour derives from the Lewis acidity of the MTO/H2O2 oxidation system, which leads to competing acid-catalysed reactions. Such a handicap may now be overcome by using additives [29]. By contrast, oxidations with isolated DMD are conducted under neutral conditions, which is highly bene®cial in the synthesis of acid-sensitive compounds. Of the two most commonly used dioxiranes, DMD and TFD, the latter is both more reactive and more selective than DMD [30].

244

W. Adam et al.

3.1 Oxidation of p Bonds (Olefins, Alkynes, Arenes)

The catalytic activity of MTO was ®rst discovered for the epoxidation of ole®ns [31]. Initial results showed drastically different yields of epoxide, which depended on the substrate. Thus, turnover numbers (TON) of 100±200 were usually obtained; however, in some cases the TON may be up to 2000 [18]. The reactivity of the MTO catalyst depends on several factors: the polarity of the solvent, the concentration of the catalyst, and the type of additive. The best reaction conditions to date are the use of 0.1 mol% MTO with 5 mol% of pyrazole as additive in tri¯uoroethanol as polar solvent, which display TONs of up to 2500 [32]. Apart from the high reactivity observed for the MTO catalyst, good selectivity may be achieved by careful control of the reaction conditions. The epoxidation of styrene serves as an example (Table 3). Thus, while MTO/ ca. 10% H2O2 shows reasonable reactivity (60% conversion after 3 h at 30 °C), the initial epoxide does not persist under the reaction conditions and subsequent Lewis-acid-catalyzed reactions lead to secondary products as a result of epoxide-ring opening, Hock cleavage, and pinacol rearrangement. These undesired transformations may be avoided by the addition of an appropriate additive [29], e.g. urea, pyridine, 3-cyanopyridine, or pyrazole. Although the presence of all of these additives affords styrene oxide in excellent product selectivity, the ef®ciency is strongly dependent on the type and the amount of additive. An increase of the amount of additive reduces the reactivity of the oxidant, while a decrease results in poorer selectivity. Table 3. Reactivity and selectivity of styrene oxidation by selected MTO-based oxidants

Oxidanta

t (h)

Convn Product distribution (%) Refs. (%) epoxide 1,2-diol aldehyde 1 aldehyde 2

MTO/H2O2 (1:100) MTO/UHP (1:100:100) MTO/H2O2/pyridine (0.5:200:12) MTO/H2O2/3-cyano-pyridine (0.5:200:10) MTO/H2O2/pyrazole (0.5:200:12)

3 19 5

60 46 70

±b ³95 ³99

37 ± ±

41 ± ±

22 ± ±

[18] [29a] [29b]

5

³99c

³99

±

±

±

[33]

5

³99

³99

±

±

±

[29b]

a

b c

All entries in CH2Cl2 except entry 1 (in t-BuOH); the ratios of catalyst to oxygen donor and/or additive is given in parentheses, with the amount of substrate being normalised to 100%. No epoxide was isolated. 85% Yield of isolated epoxide.

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

245

Two rhenium peroxo complexes [CH3ReO2(O2) and CH3ReO(O2)2 á H2O] are generated during the catalytic oxidation process and both are reported to be catalytically active. Remarkably, they possess similar reactivities towards most substrates, which include ole®ns [13]. Kinetic data are available for a series of alkyl-substituted styrenes, for which the rate constants of the two peroxo complexes are comparable. As a comparison, the rates of the DMD epoxidation with these substrates are similar [34], but TFD is generally more reactive by a factor of ca. 100 [30a]. DMD and TFD, which are the most frequently used dioxiranes, are usually employed stoichiometrically and in isolated form. The reactivity of C@C bonds towards epoxidation by the MTO-based oxidant and dioxiranes depends on several factors, most prominently the degree and pattern of substitution at the double bond, and the electronic and steric nature of these substituents. Perusal of the literature reveals that the orders of reactivity given in Fig. 3 apply for the MTO/H2O2 and DMD oxidants. The mechanistic aspects of these reactivities shall be discussed in Sect. 4. These reactivity orders are expected for electrophilic oxidants, for which electron-rich double bonds, i.e., double bonds with electron-donating substituents and/or a higher degree of substitution, are more prone to oxidation. Electronic effects of the substituents follow the order OR > R > Ar > COR > sCO2R, but steric properties operate as well. Thus, cis ole®ns are usually oxidized faster than trans ole®ns, except in the case of cis- and trans-cyclooctene, for which the latter reacts faster by a factor of 100 in the epoxidation with DMD [35]. While the MTO/UHP (UHP = urea-hydrogen-peroxide adduct) oxidation system and DMD show much similarity in their epoxidation of ole®ns (Table 4), some differences should, nonetheless, be noted. In most cases, the yields for DMD are slightly higher than with MTO/UHP; however, double bonds with electron-withdrawing groups (e.g., 2-cyclohexenone) are ef®ciently epoxidized by DMD (convn 83%), while with MTO/UHP only marginal conversion is observed (convn  5%). Additionally, hydrolytic cleavage of the epoxide occurs for particularly vulnerable epoxides (e.g., a-methylstyrene) in

Fig. 3. Epoxidation reactivity of MTO/H2O2 and DMD oxidants as a function of double-bond

substitution

246

W. Adam et al.

Table 4. Oxidation of selected ole®ns by MTO/UHP and DMD

Ole®n

a b c d e f g

MTO/UHPa Product ratio (%)

DMD Product ratio (%)

epoxide

epoxide

:

diol

b

:

5

:

diol

³95

:

5

³95

³95

:

5

³95c

:

5

68

:

32

92c

:

8

³95d

:

5e

70f

:

30e

³95

:

5

³95g

:

5

See [29a]. See [34]. See [23b]. Diastereomeric ratio (dr) = 81:19 (cis/trans); see [36]. Minor product is 2-cyclohexenone. dr = 47:53 (cis/trans); see [37]. See [15].

the case of MTO/UHP. In contrast, the chemoselectivity is poorer for DMD, as demonstrated in the oxidation of 2-cyclohexenol, from which a mixture of epoxide and enone (CH oxidation) is obtained with DMD, but only the epoxide for MTO/UHP. Fortunately, cis/trans isomerization is not a problem for either oxidant. Alkynes are much less reactive than the corresponding alkenes, and are initially oxidized by DMD and MTO/[O] to a postulated oxirene intermediate, which subsequently undergoes either further oxidation or a 1,2-H shift to a-diketones and a,b-unsaturated ketones as the main products (Table 5). Smaller amounts of other products occur through oxidation/rearrangement or oxidative cleavage. The reactivity of MTO/H2O2 towards alkynes is somewhat inferior to that of DMD; however, by far the most reactive is TFD, which gives excellent conversion of the alkynes within minutes. The product distributions obtained with DMD and TFD are similar, but for MTO/H2O2 more a-diketone is produced. The oxidation of arenes is still more dif®cult than alkynes. As expected, the oxidation of benzenes is substantially harder to achieve than that of polycyclic arenes. DMD oxidises arenes even under the standard reaction conditions (i.e.

Alkyne

a

Oxidant

t

Convn (%) Products (%)a

MTO/H2O2

2 days

69

DMD

8h

50

"

(15)

TFD

6 min

82

"

(25)

MTO/H2O2

2 days

³97

DMD

8h

>96

TFD

2 min

>96

The yields are given in parentheses

Refs. (57)

"

(0)

[38]

"

(29)

[39]

"

(49)

[39]

(43)

(11)

(24) [38]

(18)

(78)

[39]

(77)

[39]

(19)

"

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

Table 5. Oxidation of alkynes by MTO/H2O2 and DMD

247

248

W. Adam et al.

0±20 °C, acetone as solvent), whereas the reactions with MTO/H2O2 are performed in glacial acetic acid [40] and usually at elevated temperatures. The role of the acetic acid in the case of MTO/H2O2 seems to be one of increasing the electrophilicity of the peroxo complex by protonation of a peroxy oxygen atom [18]. The MTO-based oxidant and the dioxiranes show different product selectivity in the oxidation of arenes (Scheme 2), although epoxides are initially formed. In the case of DMD, these are isolable [44], while for the MTO oxidant the epoxides do not persist under the reaction conditions and rearrange in-situ to the hydroquinones. These are subsequently oxidized further to the corresponding quinones. Thus, MTO/H2O2/AcOH leads predominantly to the para-quinones. With biphenylene the ortho-quinone was obtained, while phenanthrene is overoxidized to diphenyl-2,2¢-dicarboxylic acid. The formation of undesired side-products from reactions of epoxide intermediates may be a problem (Scheme 3). The epoxides may undergo ring-opening with H2O2 to form the b-hydroxy hydroperoxide, which give water-soluble products after further oxidation. For the 3,5-di-tert-butyl derivative, the corresponding muconic anhydride is formed, while for the 2,6-dimethyl case a [4 + 2] cycloadduct is observed as side-product. In the oxidation of arenes, DMD gives epoxides in the majority of cases; exceptions are anthracene and the electron-rich 1,2,3-trimethoxy-5-methylbenzene, which give the respective para-quinones. TFD oxidizes 2-methylnaphthalene to the diepoxide (the same product as with DMD) in 66% yield after only 1 h at 0 °C [44], but phenanthrene is transformed to the corresponding ortho-quinone (9,10-quinone) in ca. 75% yield [14c]. Similar behavior is observed in the oxidation of phenols with all three oxidants, i.e., a mixture of hydroquinones and quinones (although dioxiranes have a greater propensity towards the formation of ortho-quinones than does MTO/H2O2/ AcOH) [40b, 46]. However, in the oxidation of catechol with TFD, oxidative cleavage of the benzene ring occurs and Z,Z-muconic acid is obtained in high yield [47]. Some of the quinones are of industrial interest. For example, 2-methyl-1,4naphthoquinone (Vitamin K3) is produced on an industrial scale as a supplement for animal feed [40a], while 2,3-dimethoxy-5-methyl-p-benzoquinone is a key intermediate in the synthesis of coenzyme Q [40c]. 3.2 Oxidation of Heteroatoms (S, N)

The oxidation of sul®des with MTO/H2O2 and DMD leads to the corresponding sulfoxides in high yield and excellent selectivity. If an excess of oxidant is used, the reaction may be driven towards the sulfone as the major product (Scheme 4). In the case of sulfoxidations with the MTO-based oxidant, both the monoperoxo rhenium complex CH3ReO2(O2) and the diperoxo complex CH3ReO(O2)2 á H2O are claimed to be catalytically active; however, the rate of the oxygen transfer for the monoperoxo CH3ReO2(O2) is greater by approx-

249

Scheme 2. Oxidation of arenes by MTO/H2O2 versus DMD

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

250 W. Adam et al.

Scheme 3. MTO-catalysed oxidation of phenols and methoxybenzenes

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

251

Scheme 4. Oxidation of sul®des to sulfoxides and sulfones

imately one order of magnitude than for the diperoxo rhenium complex CH3ReO(O2)2 á H2O [13] (the oxidation of thiophenes is an exception since the diperoxo complex is more active towards these substrates [48]). Kinetic data for the oxidation of para-substituted thioanisoles by DMD and the monoperoxo rhenium complex CH3ReO2(O2) show the reactivity of both to be similar [49]. The rate constants follow linear Hammett relationships with reaction constants (q) of )0.98 for MTO/H2O2, )1.00 for DMD, and )0.34 for TFD. Thus, the selectivity (sulfoxide vs sulfone) order for these electrophilic oxidants is usually TFD < MTO/H2O2 < DMD. Thiophenes are oxidized with both MTO/H2O2 and DMD to the corresponding sulfones through sulfoxide intermediates [48, 50]. In the oxidation with DMD, the thiophene 1-oxides could be kinetically stabilized by bulky substituents at the 3- and 4-positions and were isolated in crystalline form [51]. Generally these sulfoxides are dif®cult to prepare with either oxidant, as the oxidation step from the sulfoxides to the sulfones proceeds at a faster rate than the reaction of the thiophenes to the 1,1-dioxide and the sulfoxides. In addition, thiophene 1-oxide readily dimerises to Diels-Alder products. MTO/H2O2 and DMD both oxidize the sulfur in the ligand sphere of organometallic complexes [13, 52]. In fact, the (thiolato)cobalt(III) complex (en)2Co(SCH2CH2NH2)(ClO4)2 was one of the ®rst sulfur-containing compounds to be oxidized by the MTO system [13]. In the oxidation of nitrogen compounds by MTO/H2O2 and DMD, numerous products are possible. Thus, from primary amines, the corresponding nitro and nitroso compounds and oximes have been obtained, along with azo and azoxy compounds and the nitroso dimers (Scheme 5). In contrast, secondary amines (Scheme 6) are oxidized to hydroxylamines, nitrones, and in some cases nitroxides. Tertiary amines give the corresponding N-oxides cleanly. The product distribution in the MTO-catalyzed oxidation of primary amines depends on the substituents on the carbon atom a to the amine functionality and on how much of the oxygen source (H2O2) is used. In general, nitro compounds are only obtained when the a-carbon atoms are tertiary, as shown for 1-aminoadamantane in Table 6. Instead of azo compounds, further oxidation to the corresponding azoxy compounds is observed under these reaction conditions, (cf. Scheme 5). Thus, the oxidation of aniline initially leads to the corresponding hydroxylamine, which is oxidised further to PhN(OH)2, followed by the elimination of H2O to give the nitroso benzene. The latter may then either be oxidized to the nitrobenzene or undergo condensation with the aniline to produce azobenzene, which is oxidized further to azoxybenzene [63]. For benzylic amines, the oxime is selectively produced, with none of the azoxy compound or nitroso dimers [55]. This is in contrast to alkylamines (e.g., n-decylamine), for which a mixture of

252

W. Adam et al.

Scheme 5. Oxidation of primary amines with MTO/H2O2

Scheme 6. Oxidation of secondary amines with MTO/H2O2

all three of these products is obtained; however, the product ratio depends on the solvent [55]. The product distribution in the oxidation of primary amines with DMD differs from that of MTO/H2O2. Thus, due to the greater oxidation power of DMD, the nitro compounds are generally the preferred products, as illustrated for 1-aminoadamantane, aniline, and n-butylamine in Table 6. Indeed, it has been shown that both phenylhydroxylamine and nitrosobenzene are readily oxidized to nitrobenzene by DMD [54]. For a few selected substrates nitroso compounds may be isolated, when a sub-stoichiometric amount of the dioxirane is used [64]. These nitroso products readily transform to the corresponding dimers, but do not tautomerize to the oximes. The oxidation of glycosyl amines by DMD to the corresponding hydroxylamines and oximes was reported, without further reaction to the nitro sugars [65]. In the case of secondary amines, e.g., N-tert-butyl benzylamine, 1.0 equiv. of DMD cleanly afforded the hydroxylamines [56], while 2.0 equiv. led selectively to the nitrones [57]. MTO/H2O2 (UHP may also be used) [66] gave the nitrone less selectively in the case of dibenzylamine, and a larger excess (4.0 equiv.) of H2O2 was needed to convert all the hydroxylamine intermediate. For the MTO oxidant, the solvent also plays an important role in regard to the product distribution of hydroxylamine vs nitrone and, thus, much more nitrone is obtained when the reaction is conducted in protic solvents instead of chlorinated ones [55, 66a]. The oxidation of unsymmetrical secondary amines may give a mixture of regioisomeric nitrones [55].

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

253

Secondary amines with tertiary alkyl groups are oxidised by both MTO/ H2O2 and DMD in excellent yields to the corresponding nitroxides [53, 58, 67], which are used as spin labels in biochemical and medicinal investigations [68]. In this transformation, the amine is initially oxidized to the hydroxylamine, followed by further oxidation to the hydroxylamine N-oxide. This latter product is expected to cleave homolytically to give the nitroxide [58]. In contrast to the complex oxidation of primary and secondary amines, the conversion of tertiary amines is quite straightforward and leads only to N-oxides in excellent yields, e.g., quinoline and isoquinoline in Table 6. A whole range of differently functionalized tertiary aromatic amines has successfully been converted to the corresponding N-oxides with MTO/H2O2 and DMD [59, 60]. The m- and p-substituted pyridine derivatives are readily oxidized by MTO/H2O2 to the corresponding N-oxides in the presence of only 0.2±0.5 mol% MTO, regardless of the electronic nature of the substituent. In contrast, o-substituted pyridines are sluggish substrates towards oxidation with MTO/H2O2 for steric reasons and require higher amounts of catalyst (5 mol% MTO) for conversion. When DMD or TFD are used as oxidant for the transformation of tertiary aromatic amines, the in-situ-generated N-oxides are partially deoxygenated with the release of singlet oxygen (1O2) [69]. Even when a large excess of dioxirane is used, full conversion cannot be reached for some substrates. Especially the more nucleophilic heteroaromatic N-oxides are deoxygenated at comparable or even higher rates than the rate of the amine oxidation. Two of the most notable examples for such deoxygenation are 4-dimethylaminopyridine-N-oxide (Scheme 7) and 2¢,3¢,5¢-triacetyladenosine-N1-oxide [70]. Such a deoxygenation process is not known for the MTO-based oxidants. Another notable oxyfunctionalization of nitrogen compounds is the oxidation of aldehyde N,N-dimethylhydrazones to nitriles and N-methylene-

Scheme 7. Formation of singlet oxygen in the deoxygenation of heteroarene N-oxides by DMD

Table 6. Oxidation of organonitrogen compounds by MTO/H2O2 and DMD

Oxidant

T (°C)

t (h)

Products

MTO/H2O2 DMD

60 20

2 0.5

88 95

MTO/H2O2 DMD

20 20

2 0.5

70 97

MTO/H2O2

20

0.5

88

MTO/H2O2

20

0.5

38

Refs.

254

Substrate

[53] [54]

30 ±

[53] [54]

[55]

CH3(CH2)9-NH2

CH3(CH2)3-NH2

11

47

[55]

CH3(CH2)3-NH2 DMD

0.5

84

0 0

0.25 0.15

99 ±

[54]

± 96

[56] [57]

W. Adam et al.

DMD (1.0 equiv) DMD (2.0 equiv)

20

1.0 1.0

66 ±

MTO/H2O2 DMD

0 0

1 0.5

100a 98a

[53] [58]

MTO/H2O2 DMD

24 0

3 2

98 98

[59] [60]

MTO/H2O2 DMD

24 25

15 2

92 92

[59] [60]

MTO/H2O2 DMD

±50 ! 20 1 0 0.05

93 97

[61] [62]

MTO/H2O2 DMD

±50 ! 20 1 0 0.05

96a 97a

[61] [62]

No epoxide was formed.

6 91

[55] [55]

255

0 0

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

a

MTO/H2O2 (1.0 equiv) MTO/H2O2 (4.0 equiv)

256

W. Adam et al.

N-methyl N-oxide in excellent yields with both MTO/H2O2 and DMD [61, 62]. The nitriles do not react further and other functional groups (e.g., double bonds) do not affect the transformation. 3.3 Oxidation of r Bonds (CAH, SiAH)

Probably the most marked difference in the reactivity of the rhenium peroxo oxidants and dioxiranes lies in their behaviour towards oxygen insertion into CAH bonds. Little has been reported on the MTO catalysis for such a transformation, as under the normally used reaction conditions it is several orders of magnitude slower than epoxidations and does not take place [49b]. Much more drastic conditions are necessary, i.e., the use of large amounts of MTO and a large excess of the oxygen source H2O2, heating, and high concentrations, in order to achieve CAH insertions (CAH to CAOH and CAOH to C@O) with the rhenium catalyst [71]. The reaction proceeds with retention of con®guration, as shown for cis- and trans-1,2-dimethylcyclohexane and trans-decalin (Table 7). The remarkable discrimination between equatorial versus axial CAH positions in the oxidation of cis- and trans-1,2dimethylcyclohexane should be noted. When several positions are available in the substrate, the CAH insertion usually stops after one oxyfunctionalization, the multiple oxidation of adamantane is an exception. The selectivity order of the CAH bonds in unactivated alkanes is tertiary > secondary > primary, which is typical for electrophilic oxidants [73]. In the oxidation of alkanes with H2O2 catalysed by MTO and in the presence of pyrazine-2-carboxylic acid, the formation of hydroperoxide has also been observed besides alcohol and its carbonyl product [75]. The oxidation of alcohols to the corresponding carbonyl products is inef®cient. Thus, only a poor yield (19%) is achieved in the synthesis of acetophenone from (1-phenyl)ethanol (cf. Table 7). Furthermore, much lower yields of ketone/aldehyde than of alcohol are usually obtained in the oxidation of cyclohexane, cyclooctane, and n-heptane. The yields of the carbonyl compounds may be improved, when the reaction is performed in neat alcohol or in acetic acid with bromide ions added as a co-catalysts [49b]. These poor results for the MTO-based oxidant are in stark contrast to the DMD oxidation of alkanes and alcohols, and in particular the oxidation of these substrates with the highly reactive TFD (up to 700 times more reactive than DMD [73]). TFD is even capable of oxidizing unactivated hydrocarbons under very mild reaction conditions and sometimes in a matter of minutes [30a]. A prime example is the quantitative conversion of cyclohexane to cyclohexanone at )22 °C in 18 min [73]! As expected for these electrophilic oxidants, primary CAH bonds are not appreciably oxidized, and tertiary CAH bonds are preferred over secondary ones [73], although substrates with only secondary CAH bonds may be oxidized to the corresponding ketones in moderate to high yields under relatively mild conditions with TFD. In fact, TFD is so reactive that all of the alcohol intermediate is converted to the carbonyl product. The use of tri¯uoroacetic anhydride as solvent allows in-situ

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

257

258

W. Adam et al.

acetylation of the alcohols, which may subsequently be deprotected by hydrolysis and in this way product selectivity in favour of the alcohol may be achieved [76]. In addition to the excellent reactivity of TFD, this oxidant also shows higher selectivity for tertiary over secondary CAH oxyfunctionalization [77] (a notable exception is norbornene [73], which is oxidized preferentially at a secondary carbon atom) and amazing site selectivity for tertiary CAH bonds in steroidal substrates [78]. A remarkable feat of TFD is its ability to perform polyoxyfunctionalisations such as the oxidation of adamantane to adamantan-1,3,5,7-tetraol in 73% yield [77]. The relative rate data for the oxidation of para-substituted cumenes was shown to correlate linearly with the Hammett substituent constants (r) and reveals that the insertion reaction is an electrophilic process with q = )2.76 [79]. A correlation between the rate constants of a series of substituted adamantanes with the Taft steric substituent constants (r*) also gave a straight line with q* = )1.08 [79]. That alcohols are the intermediates in the oxidation of alkanes to the corresponding carbonyl compounds was con®rmed by the fact that the authentic alcohols were oxidized to the ketones/aldehydes under the reaction conditions used for the oxidation of alkanes. The second step of the reaction (CHAOH to C@O) is approximately 50 times faster than the ®rst one (CHAH to CHAOH), as shown for the oxidation of cyclohexane and cyclohexanol by TFD [73, 80]. Ethers, silyl ethers, and esters derived from secondary alcohols can also be transformed to the ketones, albeit at lower rates than the corresponding alcohols [81]. The rate constants for the oxidation of secondary alcohols (RR¢CHOH with R = Ph and R = Me, Et, n-Pr, i-Pr, t-Bu) and their methyl ethers (RR¢CHOMe) have been shown to correlate linearly with the steric bulk of the substituents, and a good Hammett correlation was observed in the case of a set of methyl ethers [81]. An important preparative application of the oxidative transformation of alcohols to ketones is the desymmetrization of 1,2- and 1,3-diols to the corresponding keto alcohols [82]. The regioselectivity depends on the electronic effects of the substituents at the carbinol positions [82d]. Thus, electron-donating groups on the phenyl substituent of 1-phenyl-propane-1,2diol lead to the preferential oxidation of the alcohol functionality next to the aromatic substituent; with electron-accepting substituents the product distribution is reversed. With optically active substrates, the transformation proceeds with retention of con®guration at the chiral center adjacent to the one undergoing oxidation to a carbonyl group [82a, 83]. The enantioselective oxidation with chiral dioxiranes leads to optically active a-hydroxy ketones [84]. Oxygen insertion into a SiAH bond (77 kcal/mol) is far easier to achieve than into a CAH (99 kcal/mol) bond, due to the fact that SiAH bonds are substantially weaker than CAH bonds [85]. In addition, for the oxidation of silanes, the net energy gain is far greater because of the formation of strong SiAO bonds. Thus, both the MTO oxidant and dioxiranes are capable of oxidizing silanes to silanols under relatively mild reaction conditions (Table 8). A disadvantage of the MTO-catalyzed oxidation of silanes to silanols compared

259

Structure, Reactivity, and Selectivity of Metal-Peroxo Complexes Versus Dioxiranes

Table 8. Oxidation of triorganosilanes

Silane

EtMe2SiH

Oxidant

MTO/85% H2O2 MTO/UHP MTO/85% H2O2/NaY n-PrMe2SiH MTO/85% H2O2 MTO/UHP DMD t-BuMe2SiH MTO/85% H2O2 MTO/UHP MTO/85% H2O2/NaY DMD PhMe2SiH MTO/85% H2O2 MTO/UHP MTO/85% H2O2/NaY DMD TFD Et3SiH MTO/85% H2O2 MTO/UHP MTO/85% H2O2/NaY TFD (+)-(a-Np) MTO/85% H2O2 PhMeSiH MTO/UHP DMD TFD a

Convn Product distribution (%) Refs. (%) Silanola Disiloxane

t

T (°C)

18 h 18 h 24 h 18 h 18 h 0.5 h 18 h 18 h 24 h 0.5 h 13 h 13 h 24 h 0.5 h 4). 93: 65-124 Tytko KH, Mehmke J, Fischer S (1999) Bonding and Charge Distribution in Isopolyoxometalate Ions and Relevant Oxides - A Bond Valence Approach. 93: 125-317 Uller E, Demleitner B, Bernt I, Saalfrank RW (2000) Synergistic Effect of Serendipity and Rational Design in Supramolecular Chemistry. 96: 149-176 Umezawa H, Takita T (1980) The Bleomycins: Antitumor Copper-Binding Antibiotics. 40: 73-99 Vahrenkamp H (1977) Recent Results in the Chemistry of Transition Metal Clusters with Organic Ligands. 32: 1-56 Valach F, Koren B, Si* P, Melnik M (1984) Crystal Structure Non-Rigidity of Central Atoms for Mn(II), Fe(II), Fe(III), Co(II), Co(III), Ni(II), Cu(I1) and Zn(I1) Complexes. 55: 101151 Valdemoro C, see Fraga S (1968) 4: 1-62 Valentine JS, see Wertz DL (2000) 97: 37-60 van Bronswyk W (1970) The Application of Nuclear Quadrupole Resonance Spectroscopy to the Study of Transition Metal Compounds. 7: 87-1 13 van de Putte P, see Reedijk J (1987) 67: 53-89 van Hove MA, see Somorjai GA (1979) 38: 1-140 van Oosterom AT, see Reedijk J (1987) 67: 53-89 Vanquickenborne LG, see Ceulemans A(1989) 71: 125-159 Vela A, see Gazquez JL (1987) 66: 79-98 Verkade JG, see Hoffmann DK (1977) 33: 57-96

322

Author Index Volumes 1-97

Vincent H, see Shannon RD (1974) 19: 1-43 Vogel E, see Keppler BK (1991) 78: 97-128 von Herigonte P (1972) Electron Correlation in the Seventies. 12: 1-47 von Zelewsky A, see Daul C (1979) 36: 129-171 Vongerichten H, see Keppler BK (1991) 78: 97-128 Wallace WE, Sankar SG, Rao VUS (1977) Field Effects in Rare-Earth Intermetallic Compounds. 33: 1-55 Wallin SA, see Hoffman BM (1991) 75: 85-108 Walton RA, see Cotton FA (1985) 62: 1-49 Warren KD, see d e n GC (1974) 19: 105-165 Warren KD, see Allen GC (1971) 9: 49-138 Warren KD, see Clack DW (1980) 39: 1-141 Warren KD (1984) Calculations of the Jahn-Teller Coupling Constants for d, Systems in Octahedral Symmetry via the Angular Overlap Model. 57: 119-145 Warren KD (1977) Ligand Field Theory of f-Orbital Sandwich Complexes. 33: 97-137 Warren KD (1976) Ligand Field Theory of Metal Sandwich Complexes. 33: 97-137 Watanabe Y , Fujii H (2000) Characterization of High-Valent 0x0-Metalloporphyrins. 97: 61-90

Watson RE, Perlman ML (1975) X-Ray Photoelectron Spectroscopy. Application to Metals and Alloys. 24: 83-132 Weakley TJR (1974) Some Aspects of the Heteropolymolybdates and Heteropolytungstates. 18: 131-176

Weichhold 0, see Adam W (2000) 97: 237-286 Weissbluth M (1967) The Physics of Hemoglobin. 2: 1-125 Wendin G (1981) Breakdown of the One-Electron Pictures in Photoelectron Spectra. 45: 1-130

Wertheim GK, see Campagna M (1976) 30: 99-140 Wertz DL, Valentine JS (2000) Nucleophilicity of Iron-Peroxo Porphyrin Complexes. 97: 37-60

Weser U (1967) Chemistry and Structure of some Borate Polyol Compounds. 2: 160-180 Weser U (1968) Reaction of some Transition Metals with Nucleic Acids and Their Constituents. 5: 41-67 Weser U (1985) Redox Reactions of Sulphur-Containing Amino-Acid Residues in Proteins and Metalloproteins, and XPS Study. 61: 145-160 Weser U (1973) Structural Aspects and Biochemical Function of Erythrocuprein. 17: 1-65 Weser U, see Abolmaali B (1998) 91: 91-190 West DC, Padhye SB, Sonawane PB (1991) Structural and Physical Correlations in the Biological Properties of Transitions Metal Heterocyclic Thiosemicarbazone and S-alkyldithiocarbazate Complexes. 76: 1-50 Westlake ACG, see Wong L-L (1997) 88: 175-208 Wetterhahn KE, see Connett PH (1983) 54: 93-124 Wilcox DE, see Solomon El (1983) 53: 1-56 Wilkie J, see Gani D (1997) 89: 133-176 Willemse J, Cras JA, Steggerda JJ, Keijzers CP (1976) Dithiocarbamates of Transition Group Elements in "Unusual" Oxidation State. 28: 83-126 Williams AF, see Gubelmann MH (1984) 55: 1-65 Williams RJP, see Frausto da Silva JJR (1976) 29: 67-121 Williams RJP, see Hill HA0 (1970) 8: 123-151 Williams RJP, see Smith DW (1970) 7: 1-45 Williams RJP, see Tam S-C (1985) 63: 103-151 Williams RJP, see Thomson AJ (1972) 11: 1-46 Williams RJP, Hale JD (1973) Professor Sir Ronald Nyholm. 15: 1 and 2 Williams RJP, Hale JD (1966) The Classification of Acceptors and Donors in Inorganic Reactions. 1 : 249-281

Author Index Volumes 1-97

323

Williams RJP (1982) The Chemistry of Lanthanide Ions in Solution and in Biological Systems. 50: 79-1 19 Wilson JA (1977) A Generalized Configuration - Dependent Band Model for Lanthanide Compounds and Conditions for Interconfiguration Fluctuations. 32: 57-91 Wilson MR (1999) Atomistic Simulations of Liquid Crystals. 94: 41-64 Winkler H, see Trautwein AX (1991) 78: 1-96 Winkler R (1972) Kinetics and Mechanism of Alkali Ion Complex Formation in Solution. 10: 1-24 Wolfbeis OS, Reisfeld R, Oehme I (1996) Sol-Gels and Chemical Sensors. 85: 51-98 Wong L-L, Westlake ACG, Nickerson DP (1997) Protein Engineering of Cytochrome P450,,,. 88: 175-208 Wood JM, Brown DG (1972) The Chemistry of Vitamin BI2 - Enzymes. 11: 47-105 Woolley RG, see Gerloch M (1981) 46: 1-46 Woolley RG (1982) Natural Optical Activity and the Molecular Hypothesis. 52: 1-35 Wiithrich K (1970) Structural Studies of Hemes and Hemoproteins by Nuclear Magnetic Resonance Spectroscopy. 8: 53-121 Xavier AV, Moura JG, Moura I (1981) Novel Structures in Iron-Sulfur Proteins. 43: 187-213 Xavier AV, see Pereira IAC (1998) 91: 65-90 Yersin H, see Gliemann G (1985) 62: 87-153 Yoko T, see Sakka S (1991) 77: 89-118 Zanchini C, see Banci L (1982) 52: 37-86 Zanello P (1992) Stereochemical Aspects Associated with the Redox Behaviour of Heterometal Carbonyl Clusters. 79: 101-214 Zanoni R, see Thiel RC (1993) 81: 1-40 Zhenyang L, see Mingos DMP (1989) 71: 1-56 Zhenyang L, see Mingos DMP (1990) 72: 73-112 Zimmerman SC, Corbin PS (2000) Heteroaromatic Modules for Self-Assembly Using Multiple Hydrogen Bonds. 96: 63-94 Zorov NB, see Golovina AP (1981) 47: 53-1 19 Zumft WG (1976) The Molecular Basis of Biological Dinitrogen Fixation. 29: 1-65

Printing (Computer to Film): Saladruck, Berlin Binding: Luderitz & Bauer, Berlin